Anda di halaman 1dari 144

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/13975

SHARE

Rock-Socketed Shafts for Highway Structure Foundations

DETAILS
136 pages | | PAPERBACK
ISBN 978-0-309-09768-0 | DOI 10.17226/13975

AUTHORS
BUY THIS BOOK

John Turner; Transportation Research Board

FIND RELATED TITLES

Visit the National Academies Press at NAP.edu and login or register to get:
Access to free PDF downloads of thousands of scientic reports
10% off the price of print titles
Email or social media notications of new titles related to your interests
Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.
Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM

NCHRP SYNTHESIS 360


Rock-Socketed Shafts
for Highway Structure
Foundations
A Synthesis of Highway Practice

CONSULTANT
JOHN TURNER
University of Wyoming
Laramie, Wyoming

S UBJECT A REAS

Bridges, Other Structures, and Hydraulics and Hydrology and Soils, Geology, and Foundations

Research Sponsored by the American Association of State Highway and Transportation Officials
in Cooperation with the Federal Highway Administration

TRANSPORTATION RESEARCH BOARD


WASHINGTON, D.C.
2006
www.TRB.org

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM

NCHRP SYNTHESIS 360

Systematic, well-designed research provides the most effective


approach to the solution of many problems facing highway
administrators and engineers. Often, highway problems are of local
interest and can best be studied by highway departments
individually or in cooperation with their state universities and
others. However, the accelerating growth of highway transportation
develops increasingly complex problems of wide interest to
highway authorities. These problems are best studied through a
coordinated program of cooperative research.
In recognition of these needs, the highway administrators of the
American Association of State Highway and Transportation
Officials initiated in 1962 an objective national highway research
program employing modern scientific techniques. This program is
supported on a continuing basis by funds from participating
member states of the Association and it receives the full cooperation
and support of the Federal Highway Administration, United States
Department of Transportation.
The Transportation Research Board of the National Academies
was requested by the Association to administer the research
program because of the Boards recognized objectivity and
understanding of modern research practices. The Board is uniquely
suited for this purpose as it maintains an extensive committee
structure from which authorities on any highway transportation
subject may be drawn; it possesses avenues of communications and
cooperation with federal, state, and local governmental agencies,
universities, and industry; its relationship to the National Research
Council is an insurance of objectivity; it maintains a full-time
research correlation staff of specialists in highway transportation
matters to bring the findings of research directly to those who are in
a position to use them.
The program is developed on the basis of research needs
identified by chief administrators of the highway and transportation
departments and by committees of AASHTO. Each year, specific
areas of research needs to be included in the program are proposed
to the National Research Council and the Board by the American
Association of State Highway and Transportation Officials.
Research projects to fulfill these needs are defined by the Board, and
qualified research agencies are selected from those that have
submitted proposals. Administration and surveillance of research
contracts are the responsibilities of the National Research Council
and the Transportation Research Board.
The needs for highway research are many, and the National
Cooperative Highway Research Program can make significant
contributions to the solution of highway transportation problems of
mutual concern to many responsible groups. The program,
however, is intended to complement rather than to substitute for or
duplicate other highway research programs.

Price $39.00
Project 20-5 (Topic 36-12)
ISSN 0547-5570
ISBN 0-309-09768-1
Library of Congress Control No. 2006925246
2006 Transportation Research Board

COPYRIGHT PERMISSION
Authors herein are responsible for the authenticity of their materials and for
obtaining written permissions from publishers or persons who own the
copyright to any previously published or copyrighted material used herein.
Cooperative Research Programs (CRP) grants permission to reproduce
material in this publication for classroom and not-for-profit purposes.
Permission is given with the understanding that none of the material will be
used to imply TRB, AASHTO, FAA, FHWA, FMCSA, FTA, or Transit
Development Corporation endorsement of a particular product, method, or
practice. It is expected that those reproducing the material in this document
for educational and not-for-profit uses will give appropriate acknowledgment
of the source of any reprinted or reproduced material. For other uses of the
material, request permission from CRP.

NOTICE
The project that is the subject of this report was a part of the National
Cooperative Highway Research Program conducted by the Transportation
Research Board with the approval of the Governing Board of the National
Research Council. Such approval reflects the Governing Boards judgment that
the program concerned is of national importance and appropriate with respect
to both the purposes and resources of the National Research Council.
The members of the technical committee selected to monitor this project and
to review this report were chosen for recognized scholarly competence and
with due consideration for the balance of disciplines appropriate to the project.
The opinions and conclusions expressed or implied are those of the research
agency that performed the research, and, while they have been accepted as
appropriate by the technical committee, they are not necessarily those of the
Transportation Research Board, the National Research Council, the American
Association of State Highway and Transportation Officials, or the Federal
Highway Administration, U.S. Department of Transportation.
Each report is reviewed and accepted for publication by the technical
committee according to procedures established and monitored by the
Transportation Research Board Executive Committee and the Governing
Board of the National Research Council.

Published reports of the

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM


are available from:

NOTE: The Transportation Research Board of the National Academies, the


National Research Council, the Federal Highway Administration, the American
Association of State Highway and Transportation Officials, and the individual
states participating in the National Cooperative Highway Research Program do
not endorse products or manufacturers. Trade or manufacturers names appear
herein solely because they are considered essential to the object of this report.

Transportation Research Board


Business Office
500 Fifth Street, NW
Washington, DC 20001
and can be ordered through the Internet at:
http://www.national-academies.org/trb/bookstore
Printed in the United States of America

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished scholars engaged in scientific and engineering research, dedicated to the furtherance of science and technology
and to their use for the general welfare. On the authority of the charter granted to it by the Congress in
1863, the Academy has a mandate that requires it to advise the federal government on scientific and technical matters. Dr. Ralph J. Cicerone is president of the National Academy of Sciences.
The National Academy of Engineering was established in 1964, under the charter of the National Academy of Sciences, as a parallel organization of outstanding engineers. It is autonomous in its administration
and in the selection of its members, sharing with the National Academy of Sciences the responsibility for
advising the federal government. The National Academy of Engineering also sponsors engineering programs
aimed at meeting national needs, encourages education and research, and recognizes the superior achievements of engineers. Dr. William A. Wulf is president of the National Academy of Engineering.
The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the
services of eminent members of appropriate professions in the examination of policy matters pertaining
to the health of the public. The Institute acts under the responsibility given to the National Academy of
Sciences by its congressional charter to be an adviser to the federal government and, on its own initiative,
to identify issues of medical care, research, and education. Dr. Harvey V. Fineberg is president of the
Institute of Medicine.
The National Research Council was organized by the National Academy of Sciences in 1916 to associate
the broad community of science and technology with the Academys purposes of furthering knowledge and
advising the federal government. Functioning in accordance with general policies determined by the Academy, the Council has become the principal operating agency of both the National Academy of Sciences
and the National Academy of Engineering in providing services to the government, the public, and the
scientific and engineering communities. The Council is administered jointly by both the Academies and
the Institute of Medicine. Dr. Ralph J. Cicerone and Dr. William A. Wulf are chair and vice chair,
respectively, of the National Research Council.
The Transportation Research Board is a division of the National Research Council, which serves the
National Academy of Sciences and the National Academy of Engineering. The Boards mission is to promote
innovation and progress in transportation through research. In an objective and interdisciplinary setting,
the Board facilitates the sharing of information on transportation practice and policy by researchers and
practitioners; stimulates research and offers research management services that promote technical
excellence; provides expert advice on transportation policy and programs; and disseminates research
results broadly and encourages their implementation. The Boards varied activities annually engage more
than 5,000 engineers, scientists, and other transportation researchers and practitioners from the public and
private sectors and academia, all of whom contribute their expertise in the public interest. The program is
supported by state transportation departments, federal agencies including the component administrations of
the U.S. Department of Transportation, and other organizations and individuals interested in the
development of transportation. www.TRB.org
www.national-academies.org

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

NCHRP COMMITTEE FOR PROJECT 20-5


CHAIR
GARY D. TAYLOR, CTE Engineers
MEMBERS
THOMAS R. BOHUSLAV, Texas DOT
DONN E. HANCHER, University of Kentucky
DWIGHT HORNE, Federal Highway Administration
YSELA LLORT, Florida DOT
WESLEY S.C. LUM, California DOT
JAMES W. MARCH, Federal Highway Administration
JOHN M. MASON, JR., Pennsylvania State University
CATHERINE NELSON, Oregon DOT
LARRY VELASQUEZ, New Mexico DOT
PAUL T. WELLS, New York State DOT
FHWA LIAISON
WILLIAM ZACCAGNINO
TRB LIAISON
STEPHEN F. MAHER

COOPERATIVE RESEARCH PROGRAM STAFF


ROBERT J. REILLY, Director, Cooperative Research Programs
CRAWFORD F. JENCKS, Manager, NCHRP
EILEEN DELANEY, Director of Publications
NCHRP SYNTHESIS STAFF
STEPHEN R. GODWIN, Director for Studies and Information
Services
JON WILLIAMS, Manager, Synthesis Studies
DONNA L. VLASAK, Senior Program Officer
DON TIPPMAN, Editor
CHERYL KEITH, Senior Secretary
TOPIC PANEL
DAN A. BROWN, Auburn University
JOHN G. DELPHIA, Texas Department of Transportation
LEO FONTAINE, Connecticut Department of Transportation
G.P. JAYAPRAKASH, Transportation Research Board
PAUL PASSE, PSI, Inc.
TOM SHANTZ, California Department of Transportation
DARIN SJOBLOM, Utah Department of Transportation
JERRY A. DIMAGGIO, Federal Highway Administration (Liaison)
BEN RIVERS, Federal Highway Administration (Liaison)

Cover photograph: Bridge with surface


exposures of foundation rock.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

FOREWORD
By Staff
Transportation
Research Board

PREFACE

Highway administrators, engineers, and researchers often face problems for which information already exists, either in documented form or as undocumented experience and practice. This information may be fragmented, scattered, and unevaluated. As a consequence,
full knowledge of what has been learned about a problem may not be brought to bear on its
solution. Costly research findings may go unused, valuable experience may be overlooked,
and due consideration may not be given to recommended practices for solving or alleviating the problem.
There is information on nearly every subject of concern to highway administrators and
engineers. Much of it derives from research or from the work of practitioners faced with
problems in their day-to-day work. To provide a systematic means for assembling and evaluating such useful information and to make it available to the entire highway community,
the American Association of State Highway and Transportation Officialsthrough the
mechanism of the National Cooperative Highway Research Programauthorized the
Transportation Research Board to undertake a continuing study. This study, NCHRP Project 20-5, Synthesis of Information Related to Highway Problems, searches out and synthesizes useful knowledge from all available sources and prepares concise, documented
reports on specific topics. Reports from this endeavor constitute an NCHRP report series,
Synthesis of Highway Practice.
This synthesis series reports on current knowledge and practice, in a compact format,
without the detailed directions usually found in handbooks or design manuals. Each report
in the series provides a compendium of the best knowledge available on those measures
found to be the most successful in resolving specific problems.

During the past 25 years, much knowledge and experience has been acquired by the engineering and construction industries on the use of rock-socketed shafts for support of transportation structures. This synthesis collected, reviewed, and organized the most salient
aspects of this knowledge and experience to present it in a form useful to foundation designers, researchers, contractors, and transportation officials. The objectives of this report were
to collect and summarize information on current practices pertaining to each step of the
design process, along with the limitations; identify emerging and promising technologies;
determine the principal challenges in advancing the state of the practice; and provide suggestions for future developments and improvements in the use and design of rock-socketed
shafts.
For this TRB synthesis report a literature review was conducted on all topics related to
drilled shaft in rock or intermediate geomaterials. A questionnaire was developed and distributed to the principal geotechnical and structural engineers of U.S. state and Canadian
provincial transportation agencies. Questions were grouped into the following categories:
use of rock-socketed shafts by the agency, evaluation of rock and intermediate geomaterials, design methods for axial loading, design methods for lateral loading, structural design,
construction, and field load and integrity testing.
John Turner, Professor of Civil and Architectural Engineering, University of Wyoming,
Laramie, collected and synthesized the information and wrote the report, under the guidance of a panel of experts in the subject area. The members of the topic panel are acknowledged on the preceding page. This synthesis is an immediately useful document that records
the practices that were acceptable within the limitations of the knowledge available at the
time of its preparation. As progress in research and practice continues, new knowledge will
be added to that now at hand.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

CONTENTS

SUMMARY

CHAPTER ONE
INTRODUCTION
Background, 3
Problem Definition, 3
Scope and Objectives, 3
Methodology, 4
Organization of Synthesis, 4
Design Process, 5

CHAPTER TWO
SITE AND GEOMATERIAL CHARACTERIZATION
Scope, 8
Site Geology, 8
Field Investigations, 10
Geologic and Index Properties of Rock, 16
Engineering Properties of Rock, 20
Intermediate Geomaterials, 29
Summary, 31

33

CHAPTER THREE
DESIGN FOR AXIAL LOADING
Scope, 33
Relationship to Geomaterial Characterization, 33
Load Transfer Behavior of Rock Sockets, 33
Capacity Under Axial Loading, 37
Axial Load-Displacement Behavior, 47
Current AASHTO Practice, 51
Summary, 52

54

CHAPTER FOUR
DESIGN FOR LATERAL LOADING
Scope, 54
Design Process, 54
Rock-Socketed Foundations for Lateral Loading, 54
Analytical Methods, 55
Structural Issues, 66
Summary, 70

71

CHAPTER FIVE
CONSTRUCTION AND FIELD TESTING
Scope, 71
Construction of Rock Sockets, 71
Field Load Testing, 74
Constructability, Inspection, and Quality Assurance, 88
Examples of Difficult Geologic Conditions, 93
Summary, 95

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

97

CHAPTER SIX
CONCLUSIONS
Site and Geomaterial Characterization, 97
Design for Axial Load, 98
Design for Lateral Load, 99
Load Testing, 99
Constructability and Inspection, 99
Research Needed to Advance State of Practice, 100

102

EQUATION SYMBOLS

104

REFERENCES

110

APPENDIX A

SURVEY RESPONDENTS

111

APPENDIX B

SURVEY QUESTIONNAIRE AND RESPONSES

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

ROCK-SOCKETED SHAFTS
FOR HIGHWAY STRUCTURE FOUNDATIONS

SUMMARY Drilled shafts are one of the few structural foundation types that can be built directly into
rock. Foundations in rock are attractive because high load carrying capacities are possible
and foundation displacements can be limited to acceptable levels more readily than through
foundations in soil. Over the past 25 years, much knowledge and experience has been gained
by the engineering and construction industries with the use of rock-socketed drilled shafts
for support of transportation structures. The goal of this synthesis is to collect, review, and
organize the most salient aspects of that knowledge and experience and to present it in a form
that is useful to foundation designers, researchers, contractors, and transportation officials.
Challenges faced by foundation designers when considering rock-socketed drilled shafts
include: (1) characterizing the nature of the rock mass or intermediate geomaterial, (2) selecting appropriate design methods for analysis of axial load carrying capacity and axial loaddeformation response, (3) analysis and design for lateral loading, and (4) assessing issues of
constructability and their influence on foundation performance and costs. Each of these issues
is considered in the synthesis within the context of the overall foundation design process as
practiced by transportation agencies.
A survey questionnaire was developed and distributed to the principal geotechnical and
structural engineers of 52 U.S. transportation agencies (including Puerto Rico and the District of Columbia) and Canadian provinces. The purpose of the survey was to define the current state of practice for rock-socketed drilled shafts. Thirty-two U.S. transportation agencies
and one Canadian provincial transportation agency responded to the questionnaire.
Innovative methods for field load testing of drilled shafts, including the Osterberg Cell
and Statnamic methods, have contributed to advances in design and construction of shafts in
rock. Load testing is shown to be an integral part of several state department of transportation programs that have led to increased use of rock-socketed drilled shafts and improved design methods. These and other load testing methods for rock-socketed shafts are reviewed.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

CHAPTER ONE

INTRODUCTION

BACKGROUND

PROBLEM DEFINITION

Highway bridges represent a large investment in the U.S.


transportation infrastructure, and structural foundations
account for a significant percentage of total bridge costs.
Current foundation engineering practice in the transportation industry represents a dramatic advancement compared
with 25 years ago. Development of this topic is illustrated
by considering that NCHRP Synthesis 42: Design of Pile
Foundations (Vesic 1977) does not mention rock-socketed
drilled shafts. At the time of its publication, NCHRP Synthesis 42 was the most comprehensive study extant on
the use of deep foundations for transportation structures.
According to DiMaggio (2004), in 1980, driven piles
accounted for more than 95% of transportation market
share, based purely on repeating previous practice. Today,
the practice is oriented toward matching the foundation
type to project conditions. This has led to a wider variety
of deep foundation types selected on the basis of subsurface conditions, structural behavior, constructability, environmental constraints, and cost. A foundation type that has
steadily increased in use over this time is the drilled shaft,
a deep foundation constructed by placing fluid concrete in
a drilled hole.

The engineering problem addressed by this synthesis is


shown in Figure 1. A drilled shaft foundation is to be designed and constructed for support of a bridge structure. Subsurface conditions may consist of soil underlain by rock.
Upper portions of the rock may be partially to highly weathered, giving these materials engineering properties that are
transitional between soil and rock, sometimes referred to as
intermediate geomaterials, or IGM. Loads to be considered
for design typically are determined by AASHTO Bridge
Design Specifications, with proper consideration of load
combinations and load factors. For foundation analysis, design loads may be resolved into vertical (P), horizontal (H),
and moment (M) components at the head of the shaft. A subsurface investigation is required to provide information on all
of the geomaterials through which the shaft must be constructed and from which the foundation will derive its resistance to the design loads. The foundation designer then must
determine the required dimensions (depth and diameter)
and structural properties of drilled shafts that will provide
adequate resistance and will limit vertical and horizontal
deformations to a level that provides adequate service performance of the bridge. Trial designs are developed and
evaluated with respect to: (1) cost, (2) performance, and
(3) constructability. A major factor in all three criteria is
whether the shaft needs to be extended into the rock or
IGM layers. Rock sockets will generally increase costs,
improve load-carrying and load-displacement performance, and make construction more challenging.

A potentially effective way to use a drilled shaft is by


bearing on, or extending into, rock. To achieve the performance and economy potentials of rock-socketed shafts, designers must be aware of the many issues that affect both
cost and performance. Drilling and excavation in rock is
generally more expensive and time consuming than in soil.
Construction of a rock socket poses challenges and difficulties that are unique and may require specialized techniques,
equipment, and experience. The first issue confronted by a
foundation designer is to determine whether a rock-socketed
foundation is necessary for bridge support. Factors to consider include the nature and magnitude of structural loads
and factors related to rock mass characteristics, including
depth to rock, rock type, rock mass engineering properties,
and constructability. The additional costs and effort of construction in rock must be offset by its benefits. The principal benefits normally are higher load-carrying capacity and
the ability to limit deformations, compared with foundations
not founded in rock. To make the appropriate cost comparisons, rock-socket design must be based on rational models
of behavior that reliably predict the capacity and loaddeformation behavior.

SCOPE AND OBJECTIVES

The overall objectives of this synthesis study are to


Collect and summarize information on current practices
pertaining to each step of the process described previously, along with their limitations and sources of
uncertainty;
Identify emerging and promising technologies in each
of these areas;
Identify the principal challenges in advancing the state
of the practice; and
Provide suggestions for future developments and improvements in the use and design of rock-socketed
shafts.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

4
P
M
H

shaft design were consulted. These include Drilled Shafts:


Construction Procedures and Design Methods by ONeill
and Reese (1999) and the AASHTO LRFD Bridge Design
Specifications (3rd ed. 2004). In addition, a draft version of
Section 10, Foundations, of the 2006 Interim AASHTO
LRFD Bridge Design Specifications was reviewed.
A questionnaire was developed and sent to the principal
geotechnical and structural engineers of 52 U.S. transportation
agencies (including Puerto Rico and the District of Columbia)
and the Canadian provinces. The primary purpose of the survey
was to define the current state of practice for rock-socketed
shafts. Questions were grouped into the following categories:

SOIL

WEATHERED ROCK OR IGM

ROCK

FIGURE 1 Rock-socketed shaft designed for highway bridge


structure.

The major challenges faced by U.S. transportation agencies


in the use of rock-socketed drilled shafts for highway bridges
were identified by NCHRP Topic Panel 36-12 as follows:
The first challenge is characterizing the nature of the
rock or IGM. By its nature, rock and IGM are highly
variable and difficult to characterize for engineering
purposes. To effectively design drilled shafts in rock and
IGM, engineers must accommodate high levels of uncertainty. Issues to be addressed include quantifying
material characteristics, rock mass behavior, and appropriate application of laboratory and field test methods.
The second challenge is determination of the axial load
capacity of rock-socketed shafts. Rock-socketed shafts
resist axial load in both side shear and end bearing. Designers need well-documented methods for assessing
side shear and end bearing. Different methods are appropriate for different types of geology. There are many
issues related to characterizing the rock and construction that affect design for axial loading.
The third challenge is analysis and design of rocksocketed shafts under lateral loading. It has been a customary practice to adopt the techniques developed for
laterally loaded piles in soil to solve the problem of
rock-socketed shafts under lateral loading. There exist
several analysis and design methods specifically for
rock-socketed shafts under lateral loading; their application in practice remains limited.
METHODOLOGY

A literature review was conducted on all topics related to


drilled shafts in rock or IGM. To assess current practice, the
primary manuals used by transportation engineers for drilled

Use of rock-socketed shafts by the agency,


Evaluation of rock and IGM properties,
Design methods for axial loading,
Design methods for lateral loading,
Structural design,
Construction, and
Field load and integrity testing.

Thirty-two U.S. and one Canadian provincial transportation agencies responded to the questionnaire, completely or
in part. A list of responding agencies and a summary of responses to the questions are given in Appendix A. The questionnaire was also sent to several consulting firms and drilled
shaft contractors. Two contractors responded to the survey.
Based on responses to the questionnaire, selected state
agency personnel and contractors were interviewed.

ORGANIZATION OF SYNTHESIS

The synthesis is presented in six chapters and two appendixes.


Chapter one defines the problem, objectives, scope, and
methodology of the study. This chapter also provides an
overview of the foundation design process used by state department of transportation (DOT) agencies. This overview
provides a framework for understanding the interrelationships
between site characterization, material property evaluation,
geotechnical and structural design, load testing, and construction of rock-socketed shafts. Each of these topics is considered in subsequent chapters. Chapter two reviews methods
of site characterization and material property evaluation that
are applicable to rock-socketed shafts. Chapter three is a compilation and critical review of methods used for analysis and
design of rock sockets for axial loading. Chapter four reviews
and summarizes analysis methods for rock sockets under lateral and moment loading and discusses structural design issues relevant to rock sockets. Chapter five provides an
overview of current technologies for rock-socket construction
and considers some of the construction issues identified by
the survey. This chapter also covers field load testing of rocksocketed shafts and the role of load testing within the context
of state DOT foundation engineering programs. Chapter six

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

is a summary of the principal findings of this study and presents steps that can lead to more effective use, design, and
construction of rock sockets for bridge foundations. In each
chapter, significant findings derived from the survey are identified and discussed. Appendix A provides a list of survey respondents and Appendix B presents the questionnaire and a
compilation of the responses to each question.

Structures Office (Bridge). Information provided includes a


brief description of the anticipated site conditions, conceptual
foundation types considered to be feasible, and conceptual
evaluation of potential geotechnical hazards such as liquefaction. The purpose of these recommendations is to provide
sufficient geotechnical information to allow a bridge preliminary plan to be produced.

DESIGN PROCESS

Develop Site Data and Preliminary Plan

Structural foundation design within state DOTs is typically a


joint effort between the structural and geotechnical divisions.
The geotechnical group may include engineering geologists
and both groups may operate under the supervision of a chief
bridge engineer. As a starting point, consider Figure 2, which
shows a flow chart of the overall foundation design process.
The chart is from the Washington State DOT Geotechnical
Design Manual (2005). Based on responses to Question 4 of
the survey questionnaire (Appendix B) and interviews with
DOT personnel, Figure 2 typifies the process followed by
many states. A summary of each step, also based on the
Washington State DOT manual (2005), is as follows.

Bridge obtains site data from the regional office and develops a preliminary bridge plan (or other structure) adequate
for GD to locate borings in preparation for final design of the
structure. Bridge would also provide the following information to GD to support development of the preliminary foundation design:

Conceptual Bridge Foundation Design

An informal communication/report is produced by the Geotechnical Division (GD) at the request of the Bridge and
Bridge and Structures Office
(Bridge) requests conceptual
foundation recommendations
from Geotechnical Division
(GD)

GD provides conceptual
foundation recommendations
to Bridge

Anticipated structure type and magnitudes of tolerable


settlement (total and differential).
At abutments, the approximate maximum elevation feasible for the top of the foundation.
For interior piers, the number of columns anticipated
and, if there will be single foundation elements for each
column or if one foundation element will support multiple columns.
At stream crossings, the depth of scour anticipated, if
known. Typically, GD will pursue this issue with the
Hydraulics Office.
Known constraints that would affect the foundations in
terms of type, location, or size, or any known constraints that would affect the assumptions made to determine the nominal resistance of the foundation (e.g.,
utilities that must remain, construction staging needs,
excavation, shoring and falsework needs, and other
constructability issues).
Preliminary Foundation Design

Bridge obtains site data,


develops draft preliminary
plan, and provides initial
foundation needs input to GD

GD provides preliminary
foundation design
recommendations
iterate

Bridge performs structural


analysis and modeling and
provides feedback to GD
regarding foundation loads,
type, size, depth, and
configuration needed for
structural purposes

GD performs final
geotechnical design and
provides final geotechnical
report for the structure

Bridge performs final structural


modeling and develops final
PS&E for structure

FIGURE 2 Design process for Load and Resistance Factor


Design (Washington State DOT 2005). PS&E = Plans,
Specifications, & Estimates.

A memorandum is produced by GD at the request of Bridge


that provides geotechnical data adequate to conduct structural analysis and modeling for all load groups to be considered. The geotechnical data are preliminary and not in final
form for publication and transmittal to potential bidders. At
this stage, foundation recommendations are subject to
change, depending on the results of structural analysis and
modeling and the effect that modeling and analysis has on
foundation types, locations, sizes, and depths, as well as any
design assumptions made by the geotechnical designer. Preliminary foundation recommendations may also be subject
to change based on construction staging needs and other
constructability issues discovered during this phase. Geotechnical work conducted during this stage typically includes completion of the field exploration program to the final PS&E
level (Plans, Specifications, & Estimates), development of
foundation types and feasible capacities, foundation depths

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

needed, p-y curve data and soil spring data for seismic modeling, seismic site characterization and estimated ground
acceleration, and recommendations to address known constructability issues. A description of subsurface conditions
and a preliminary subsurface profile would also be provided
at this stage; however, detailed boring logs and laboratory
test data would usually not be provided.

Anticipated foundation loads (including load factors


and load groups used),
Foundation size/diameter and depth required to meet
structural needs,
Foundation details that could affect the geotechnical
design of the foundations, and
Size and configuration of deep foundation groups.

Structural Analysis and Modeling

Final Foundation Design

Bridge uses the preliminary foundation design recommendations provided by GD to perform structural modeling of the
foundation system and superstructure. Through this modeling, Bridge determines and distributes the loads within the
structure for all appropriate load cases, factors the loads as
appropriate, and sizes the foundations using foundation nominal resistances and resistance factors provided by GD. Constructability and construction staging needs continue to be
investigated during this phase. Bridge provides the following
feedback to GD to allow them to check their preliminary
foundation design and produce the Final Geotechnical Report for the structure:

This design step results in a formal report produced by GD


that provides final geotechnical recommendations for the
subject structure. This report includes all geotechnical data
obtained at the site, including final boring logs, subsurface
profiles, laboratory test data, all final foundation recommendations, and final constructability recommendations for the
structure. At this time, GD checks the preliminary foundation
design in consideration of the structural foundation design
results determined by Bridge, and makes modifications as
needed to accommodate the structural design needs provided
by Bridge. Some state DOTs may also make this report available to potential bidders.

FIGURE 3 Design and construction process for drilled shaft foundations (adapted from Paikowsky et al. 2004a). QA/QC =
quality assurance/quality control; NDT/NDE = nondestructive testing/evaluations.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

7
Final Structural Modeling and Development
of Plans, Specifications, and Estimates

Bridge makes the required adjustments to the structural model


to accommodate changes in the geotechnical foundation
recommendations as transmitted in the final geotechnical
report. From this, the bridge design and final PS&E are
completed. A similar design process is recommended if a consultant or designbuilder is performing one or both design
functions.

Design Process in Relation to the Synthesis

Based on the process described previously and followed by


most state DOTs, Figure 3 is a flowchart of the design and
construction process for drilled shaft foundations that provides a framework for the topics addressed by this synthesis.
In each subsequent chapter, the topics being covered are considered within the context of the overall process as shown in
Figure 3. This includes site investigation, geomaterial property evaluation, and design for axial and lateral loading.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

CHAPTER TWO

SITE AND GEOMATERIAL CHARACTERIZATION

SCOPE

SITE GEOLOGY

This chapter describes site investigation methods, classification systems for intact rock and rock masses, and field and
laboratory tests used to determine rock engineering properties. The focus is limited to information relevant to the design
and construction of rock-socketed drilled shafts. Several
references are available that provide guidance on strategies
and methods of site characterization and material property
evaluation for geotechnical practice, with a focus on transportation facilities. These include the FHWA Manual on
Subsurface Investigations (Mayne et al. 2001), Evaluation
of Soil and Rock Properties, Geotechnical Engineering
Circular No. 5 (Sabatini et al. 2002), and the AASHTO Manual on Subsurface Investigations (1988). In addition, the U.S.
Army Corps of Engineers has published several manuals relevant to this topic (Rock Testing Handbook 1993; Rock
Foundations 1994; Geotechnical Investigations 2001).

Understanding the geologic environment provides information used to plan the more detailed, subsequent phases of exploration. Site geology refers to the physiography, surficial
geology, and bedrock geology of the site. The starting point
is a thorough survey of existing information. In many cases,
existing data will enable identification of geologic features
that will determine the feasibility of rock-socketed foundations or will have a major impact on their design or construction. The amount and quality of information gathered
can then be used to establish the type and extent of additional
data that will be required. General knowledge of the site
geology is required in the first phase of the design process
outlined in chapter one, Conceptual Bridge Foundation
Design, to establish anticipated site conditions, feasibility of
rock sockets, and conceptual evaluation of potential geotechnical hazards.

The purpose of site characterization is to obtain the information required to develop a model of the site geology and to
establish the required engineering properties of the geomaterials. The information obtained is used for two general purposes:
(1) analysis of capacity and load-deformation response, which
determines the foundation overall design; and (2) construction
feasibility, costs, and planning. Once the site for a bridge or
other transportation structure has been established, all aspects
of the site and material characterization program are focused
on the soil and rock conditions as they exist at that site.
Geologic conditions and rock mass characteristics can exhibit
such a wide degree of variability that it is not possible to establish a single standardized approach. The scope of the program
is determined by the level of complexity of the site geology,
foundation loading characteristics, size, configuration, and
structural performance of the bridge, acceptable levels of risk,
experience of the agency, and other factors. Some of the information needed to establish the scope of site characterization
may only be known following a preliminary study of the site.

Sources of existing data include: geologic and topographic


maps, publications, computer databases, aerial photographs,
and consultation with other professionals. Many references
are available that provide detailed information on sources and
applications of existing data to geotechnical site characterization (e.g., Mayne et al. 2001). A detailed treatment of the topic
is beyond the scope of this report and only the general aspects
of such data sources will be summarized.

Rock and IGM exhibit behaviors that are unique and


require special techniques for application to engineering
problems. Two aspects of rock behavior that are paramount
are: (1) natural rock masses may exhibit a high degree of
variability and (2) properties of a rock mass are determined
by the combined properties of intact rock and naturally
occurring discontinuities, such as joints, bedding planes,
faults, and other structural features.

Geologic maps are used to transmit information about geologic features at or near the earths surface. Maps are prepared at various scales and for a variety of purposes (Varnes
1974). A geologic map may be prepared to depict the general
geology of a large region, for example bedrock geology of an
entire state, or it may cover a relatively small area and contain detailed information about specific geologic features, for
example engineering geology of a single quadrangle. A good
starting point is the geologic map of the state. These maps are
produced at a scale that makes it possible to identify the
underlying bedrock formations in a general area. Often this
is sufficient to know immediately whether a bridge is located
where bedrock conditions are favorable or unfavorable for
foundations in rock, or even whether bedrock exists at reasonable depth. Most state DOT geotechnical engineers and
geologists with experience have familiarity with the geology
of their state and incorporate this step unconsciously. The
next logical step is to determine if more detailed geologic
maps or reports are available for the particular area in which

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

the bridge is located. Sources of such maps and publications


include U.S. Geological Survey and state Geological Surveys, university libraries, and Soil Conservation Service. The
use of Internet search engines has added a powerful tool for
locating such information and most governmental geologic
publications can now be identified and obtained on-line.
Detailed geologic maps normally provide useful information
on characteristics of bedrock and, in some cases surficial,
geology relevant to foundation engineering. These maps provide descriptions of rocks in terms of lithology (rock type,
mineralogy, and genesis), age, and structure (strike and dip
of sedimentary rocks). In addition, major structural features
are identified, such as faults, folds, and contacts between
rock units (formations or members). Geologic maps prepared
specifically for engineering purposes may include data on
discontinuity patterns and characteristics, rock material
strength, Rock Mass Ratings (RMRs), groundwater conditions, and depth to bedrock (Radbruch-Hall et al. 1987).
Many will identify geologic hazards such as swelling soils
or rock, landslides, corrosion potential, karst, abandoned
mines, and other information of value. If engineering geologic maps are available they are an essential tool that
should be used.
The most practical aerial photographs for geotechnical
purposes are black and white photographs taken with stereo
overlap and with panchromatic film, from heights of between
500 m and 3,000 m, at scales of about 1:10,000 to 1:30,000.
The higher level photographs provide a resolution most useful for larger-scale features such as topography, geology, and
landform analysis, whereas the lower-level photographs provide more detail on geologic structure. Landslides and debris
flows, major faults, bedding planes, continuous joint sets,
rock outcrops, and surface water are some of the features
that can be identified and are relevant to the siting of bridge
structures.
A potentially valuable source of existing data may be
consultation with other geoprofessionals with design or construction experience in the same rock units. Geotechnical
engineers, geologists, groundwater hydrologists, contractors, mining company personnel, well drillers, etc., may be
able to provide geotechnical engineering reports from
nearby projects, photographic documentation of excavations
or other construction works, and unpublished reports or testing data. In addition, such individuals are often willing to
share relevant experience. Bedian (2004) describes a case
history in which experience at an adjacent site was used to
develop a value engineer proposal for the design of rocksocketed foundations for a high rise building.
The geotechnical literature contains many useful papers
describing design, construction, and/or load testing of rocksocketed drilled shafts in which the focus is on a particular
type of rock or a specific formation. For example, Hassan
and ONeill (1997) present correlations for side resistance

of shafts in the Eagle Ford Shale, a rock unit commonly


encountered in north-central Texas, most notably in the
Dallas area. Results of load tests on drilled shafts in mica
schist of the Wissahickon Formation, commonly encountered in Philadelphia and other parts of eastern Pennsylvania,
are given in Koutsoftas (1981) and Yang et al. (2004). Turner
et al. (1993) and Abu-Hejleh et al. (2003) consider side resistance from load tests on shafts socketed into Pierre and
Denver Formation shales. McVay et al. (1992) present a thorough study on the design of shafts in Florida limestone.
Numerous other examples could be cited. Whenever such
publications are available they should be used as a source of
background information during the planning phase of any
project where the same rock units are present. Results of load
tests at different locations, but in the same rock unit, cannot
be applied without judgment and site-specific considerations, but they do provide a framework for considering
design issues and may provide insight on expected performance. Similarly, publications describing construction challenges in certain geologic environments and strategies for
addressing them can be useful. Schwartz (1987) described
construction problems and recommended solutions for rocksocketed piers in Piedmont formations in the Atlanta area.
Brown (1990) identified problems involved in construction
of drilled shafts in the karstic limestone of northern Alabama
and suggests methods and approaches that have been successful for dealing with such challenges. A literature review
often is all that is necessary to locate this type of useful
information.
Where bedrock is exposed in surface outcrops or excavations, field mapping is an essential step to obtaining information about rock mass characteristics relevant to design
and construction of foundations. A site visit is recommended for reconnaissance and field mapping following a
review of existing information. A competent engineering
geologist or geotechnical engineer can make and record observations and measurements on rock exposures that may
complement, or in some cases exceed, the information obtained from borings and core sampling. Rock type, hardness, composition, degree of weathering, orientation and
characteristics of discontinuities, and other features of a
rock mass may be readily assessed in outcrops or road cuts.
Guidance on detailed geologic mapping of rock for engineering purposes is given in Murphy (1985), Rock
Slopes . . . (1989), and ASTM D4879 (Annual Book . . .
2000). Photography of the rock mass can aid engineers and
contractors in evaluating potential problems associated with
a particular rock unit. The major limitation lies in whether
the surface exposure is representative of the rock mass at a
depth corresponding to foundation support. When rock coring and surface mapping demonstrate that surface exposures
are representative, the surface exposures should be exploited for information. Figure 4 shows a bridge site where
mapping of rock exposures could provide much of the relevant data for design of foundations.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

10

NCHRP Synthesis 357: Use of Geophysics for Transportation Projects (Sirles 2006) provides a comprehensive
overview of the topic and additional survey data relevant to
this study. Table 1 identifies the primary and secondary
methods used to investigate selected subsurface objectives.
The table is an abridged version from the Sirles report (2006)
in which only objectives pertaining to foundation investigations are included. The survey of transportation agencies
for this project identified seismic as the most widely used
geophysical methods and mapping rock as the most widely
used application of geophysics. Mapping karst or other voids
was also identified as a major objective.
FIGURE 4 Bridge site with surface exposures of foundation
rock.

FIELD INVESTIGATIONS

Field methods for characterization of rock include geophysical methods, rock core drilling, and in situ testing. These
activities normally are carried out during the Preliminary
Foundation Design phase of the design process as described
in chapter one, and would be used to provide a description
of subsurface conditions and a preliminary subsurface profile. The detailed results of field investigations, including
detailed boring logs, in situ testing results, and interpretation, would be included in the final geotechnical report prepared during the Final Foundation Design phase of Figure 2.
Geophysical Methods

Geophysical methods, in conjunction with borings, can provide useful information in areas underlain by rock. The most
common application of geophysics is to determine depth to
bedrock. When correlated with data from borings, geophysical methods provide depth to bedrock information over a
large area, eliminating some of the uncertainty associated
with interpolations of bedrock depths for locations between
borings.
Geophysical methods are based on measuring the transmission of electromagnetic or mechanical waves through the
ground. Signal transmission is affected by differences in the
physical properties of geomaterials. By transmitting electromagnetic or seismic signals and measuring their arrival at
other locations, changes in material properties can be located.
In some cases, the material properties can also be quantified.
For foundation site characterization, geophysical methods can
be placed into two general categories, those conducted from
the ground surface (noninvasive) and those conducted in
boreholes (invasive). When grouped according to method, the
six major categories are: seismic, electromagnetic, electrical,
magnetic, radar, and gravity. Basic descriptions of geophysical methods and their application to geotechnical engineering
are given by the U.S. Army Corps of Engineers (Geophysical Exploration . . . . 1995) and Mayne et al. (2001).

Results of the survey for this study are consistent with


those of Sirles (2006). The most frequently applied method
is seismic refraction, which is based on measuring the travel
time of compressional waves through the subsurface. Upon
striking a boundary between two media of different properties the direction of travel is changed (refraction). This
change in direction is used to deduce the subsurface profile.
Figure 5a illustrates the basic idea for a simple two-layer
profile in which soil of lower seismic velocity (Vp1) overlies
rock of higher seismic velocity (Vp2). A plot of distance from
the source versus travel time (Figure 5b) exhibits a clear
change in slope corresponding to the depth of the interface.
The equipment consists of a shock wave source (typically a
hammer striking a steel plate), a series of geophones to measure seismic wave arrival, and a seismograph with oscilloscope. The seismograph records the impact and geophone
signals in a timed sequence and stores the data digitally. The
technique is rapid, accurate, and relatively economical when
applied correctly. The interpretation theory is relatively
straightforward and equipment is readily available. The most
significant limitations are that it is incapable of detecting
material of lower velocity (lower density) underlying higher
velocity (higher density) and that thin layers sometimes are
not detectable. For these reasons, it is important not to rely
exclusively on seismic refraction, but to verify depth to rock
in several borings and correlate the seismic refraction signals
to the boring results. Seismic velocity, as determined from
seismic refraction measurements, can be correlated to smallstrain dynamic modulus of soil and rock by the following
relationships:
Ed = 2 (1 + vd ) Vs2
Ed =

(1 2 v ) (1 + vd ) 2
Vp
(1 vd )

(1)
(2)

in which Ed = small-strain dynamic modulus, vd = smallstrain dynamic Poissons ratio, = mass density, Vs = shear
wave velocity, and Vp = compressional wave velocity. Eqs. 1
and 2 are based on the assumption that the rock mass is a
homogeneous, isotropic, elastic solid. Because most rock
masses depart significantly from this assumption, elastic
modulus values calculated from seismic wave velocities are
normally larger than values measured in static field load

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

11
TABLE 1
GEOPHYSICAL METHODS AND APPLICATIONS (after Sirles 2006)
Techniques

Ground Penetrating Radar

Magnetics

Other

Gravity

Electrical Resistivity Tomography/P

EM61TimeDomain Metal Detector

EM34Terrain Conductivity

Electrical Resistivity/P

P
P

Electrical

TimeDomain EM Soundings

EM31Terrain Conductivity

Shear Wave

P
P

Seismic Tomography

Seismic Reflection

Investigation Objectives
Bedrock Depth
Rippability
Lateral & Vertical
Variation in Rock or Soil
Strength
Location of Faults and
Fracture Zones
Karst Features

Seismic Refraction

Methods

Electromagnetic

Surface Wave (SASW, MASW, &


Passive)

Seismic

S
P

S
P

Notes: P = primary; S = secondary; blank = techniques should not be used; EM = electromagnetic; SASW = spectral analysis of surface
waves; MASW = multi-channel analysis of surface waves.

tests, such as plate bearing or pressure chamber tests. Alternatively, a method that correlates rock mass modulus to shear
wave frequency has been shown to provide a reasonable firstorder estimate of modulus. Figure 6 shows the relationship
between in situ modulus and shear wave frequency using a
hammer seismograph, as described by Bieniawski (1978).
The data can be fit to a straight line by
EM = 0.054f 9.2

(3)

where EM = rock mass static modulus (GPa) and f = shear


wave frequency (hertz) from the hammer blow received at
distances of up to 30 m on a rock surface.
Resistivity is a fundamental electrical property of geomaterials that varies with material type and water content. To
measure resistivity from the ground surface (Figure 7), electrical current is induced through two current electrodes (C1
and C2), while change in voltage is measured by two potential electrodes (P1 and P2). Apparent electrical resistivity is
then calculated as a function of the measured voltage difference, the induced current, and spacing between electrodes.
Two techniques are used. In a sounding survey, the centerline of the electrodes is fixed while the spacing of the electrodes is increased for successive measurements. The depth
of material subjected to current increases with increasing
electrode spacing. Therefore, changes in measured apparent
resistivity with increasing electrode spacing are indicative
of a change in material at depth. In this way, variations in

material properties with depth (layering) can be determined.


The second method is a profiling survey in which the electrode spacing is fixed but the electrode group is moved
horizontally along a line (profile) between measurements.
Changes in measured apparent resistivity are used to deduce
lateral variations in material type. Electrical resistivity
methods are inexpensive and best used to complement seismic refraction surveys and borings. The technique has advantages for identifying soft materials in between borings.
Limitations are that lateral changes in apparent resistivity
can be interpreted incorrectly as depth related. For this and
other reasons, depth determinations can be in error, which
is why it is important to use resistivity surveys in conjunction with other methods.
The use of multi-electrode resistivity arrays shows promise
for detecting detailed subsurface profiles in karst terranes, one
of the most difficult geologic environments for rock-socketed
foundations. Dunscomb and Rehwoldt (1999) showed that
two-dimensional (2-D) profiling using multi-electrode arrays
provides reasonable resolution for imaging features such as
pinnacled bedrock surfaces, overhanging rock ledges, fracture zones, and voids within the rock mass and in the soil
overburden. Hiltunen and Roth (2004) present the results of
multiple-electrode resistivity surveys at two bridge sites on
I-99 in Pennsylvania. The resistivity profiles were compared with data from geotechnical borings. Both sites are
located in karst underlain by either dolomite or limestone.
The resistivity profiles provided a very good match to the

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

12

(a)

FIGURE 6 Rock mass modulus versus shear wave frequency


by Bieniawski (Goodman 1980).

(b)

FIGURE 5 Seismic refraction method (Mayne et al. 2001):


(a) field setup and procedures; (b) data reduction for depth to
hard layer.

stratigraphy observed in borings, particularly for top-ofrock profile. Figure 8 shows a resistivity tomogram at one
of the bridge pier sites, in which the top-of-rock profile is
well-defined by the dark layer. Inclusions of rock in the
overlying soil are also clearly defined. This technology
should be considered for any site where a rock surface profile is required and would provide valuable information for
both design and construction of rock-socketed foundations. Table 1 identifies electrical resistivity tomography
profiling as a primary method for investigating karstic conditions and as a secondary method for measuring depth to
bedrock.
Other geophysical methods have potential for rock sites,
but have yet to be exploited specifically for applications to
foundations in rock. These include downhole and crosshole
seismic methods. Downhole seismic p-wave is based on measuring arrival times in boreholes of seismic waves generated

at the ground surface. Crosshole seismic involves measuring travel times of seismic waves between boreholes. Both
methods provide depth to rock, and s-wave velocities,
dynamic shear modulus, small-strain Youngs modulus, and
Poissons ratio. Crosshole tomography is based on computer
analysis of crosshole seismic or resistivity data to produce a
3-dimensional (3-D) representation of subsurface conditions.
These techniques are more expensive and require specialized
expertise for data interpretation, but may be cost-effective for
large structures where the detailed information enables a
more cost-effective design or eliminates uncertainty that may
otherwise lead to construction cost overruns.
All geophysical methods have limitations associated with
the underlying physics, the equipment, and the individuals
running the test and providing interpretation of the data. The
study by Sirles (2006) includes several informative case histories from state DOTs of both successful and unsuccessful
projects. The single case history related to a bridge foundation investigation is one of a failure to provide accurate

Current meter

Battery

Volt meter
P1
C1

Spacing, A

P2
V
Spacing, A

Spacing, A

FIGURE 7 Field configuration for resistivity test.

Copyright National Academy of Sciences. All rights reserved.

C2

Rock-Socketed Shafts for Highway Structure Foundations

13
Resistivity Test #7
West

Depth (feet)

East

-10

-20
10

20

30

40

50

60

70

80

90

100

Distance (feet)

200

100

50

25

12

Resistivity (Ohm-feet)

FIGURE 8 Resistivity tomogram at Pennsylvania bridge site in


karst (Hiltunen and Roth 2004).

depths to bedrock in a river channel using both seismic


refraction and an electrical resistivity sounding survey. Reasons cited for the failure include loss of geophones owing
to running water and ice, instrumentation malfunctions, excessive background noise, differences of opinion between
consultants on data interpretation, and discrepancies between
top of rock from geophysical results and borings. Although
this is not believed to be a typical case, it demonstrates some
real world lessons.
Additional findings by Sirles (2006) are that in-house
geoscientists and engineers do not understand the value, the
benefit, or the science of geophysics for their projects. However, several factors point to geophysics becoming more
widely accepted and implemented as a tool in the transportation industry. These include a manual published by FHWA
and available on-line (http://www.cflhd.gov/geotechnical),
additional programs aimed at training of agency personnel,
and increasing levels of experience.
Borings

Borings provide the most direct evidence of subsurface conditions at a specific site. They furnish detailed information
on stratigraphy and samples of soil and rock from which
engineering properties are determined. Borings also provide
the means for conducting in situ tests, installation of instrumentation, and observing groundwater conditions. Conventional soil boring and testing equipment is used to drill
through overlying soil deposits and to determine depth to
bedrock. Once encountered, the most widely used technique
for investigating rock for the purpose of foundation design
is core drilling. Samples are obtained for rock classification
and determining rock properties important to both design and
construction. A core sample can be examined physically and
tested, providing information that is hard to obtain by any
other methods.
Rock core drilling is accomplished using rotary drill
equipment, usually the same truck- or skid-mounted rigs
used for soil drilling and sampling. A hollow coring tube
equipped with a diamond or tungstencarbide cutting bit is

rotated and forced downward to form an annular ring while


preserving a central rock core. Standard core barrel lengths
are 1.5 m and 3 m (5 ft and 10 ft). Fluid, usually water but
possibly drilling mud, is circulated for cooling at the cutting
interface and removal of cuttings. Selecting the proper tools
and equipment to match the conditions and the expertise of
an experienced drill crew are essential elements of a successful core drilling operation. Once rock is encountered,
coring normally is continuous to the bottom of the hole.
Where the rock being sampled is deep, wire line drilling,
in which the core barrel is retrieved through the drill stem,
eliminates the need to remove and reinsert the entire drill
stem and can save considerable time. If sampling is not continuous, drilling in between core samples can be accomplished using solid bits.
Rock coring bits and barrels are available in standardized
sizes and notations. Important considerations in core barrel
selection are: (1) core recovery and (2) the ability to determine the orientation of rock mass structural features relative
to the core. Core recovery is most important in highly fractured and weak rock layers, because these zones are typically
critical for evaluation of foundationrock load transfer.
For sampling of competent rock, bits and core barrels that
provide a minimum of 50-mm-diameter (nominal) core are
adequate for providing samples required for index tests, rock
quality designation (RQD), laboratory specimens for
strength testing, and evaluating the conditions of discontinuities. For example, NWM (formerly NX) diamond bit and
rock core equipment drills a 76-mm (3-in.) diameter hole and
provides a 54-mm (2.125-in.) diameter rock core. When weak,
soft, or highly fractured rock is present, it may be necessary
to use larger diameter bits and core barrels to improve core
recovery and to obtain samples from which laboratory strength
specimens can be prepared. Coring tools up to 150 mm
(6 in.) in diameter are used. A highly recommended practice
for best core recovery is to use triple-tube core barrels. The
inner sampling tube does not rotate during drilling and is
removed by pushing instead of hammering; features that
minimize disturbance. Thorough descriptions of coring
equipment and techniques are given in Acker (1974),
AASHTO Manual on Subsurface Investigations (1988),
Mayne et al. (2001), and U.S. Army Corps of Engineers
(Geotechnical Investigations 2001).
Steeply dipping or near-vertical bedding or jointing may
go undetected in holes drilled vertically (Terzaghi 1965).
Such features can significantly influence the strength and deformability of rock foundations. Inclined (nonvertical)
drilling provides the opportunity to detect the orientation and
characteristics of near-vertical features. Oriented core refers
to any method that provides a way to determine the geometrical orientation of planar structural features, such as bedding, joints, fractures, etc., with respect to the geometrical
orientation of the core. One approach is to mark the core with
a special engraving tool so that the orientation of the discontinuity relative to the core is preserved and the orientation of

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

14

the discontinuity (strike and dip) can be determined accurately (Goodman 1976). A method used with wire line
drilling involves making an impression of the core in clay.
The combination of inclined and oriented coring techniques
can provide an effective tool for characterizing orientation of
discontinuities in complexly fractured rock masses. Rock
core orienting methods are covered in more detail in the
AASHTO Manual on Subsurface Investigations (1988) and
are also reviewed and compared with borehole televiewer
methods by Eliassen et al. (2005).
Depth and Spacing of Boreholes

ONeill and Reese (1999) recommend the number of borings


to be made per drilled shaft location at bridge sites when the
material to be excavated is unclassified (Table 2). Unclassified means the contractor is paid by the unit of excavation
depth (meters or feet) regardless of the material encountered.
For rock sites, these recommendations should be considered
a minimum. If possible, it is recommended to locate one
boring at every rock-socketed shaft. In practice, this is not
always possible and factors such as experience, site access,
degree of subsurface variability, geology, and importance
of the structure will be considered. If materials are classified for payment purposes, it becomes more important to
locate a boring at every drilled shaft location for the purpose
of making accurate cost estimates and for contractors to
base their bids on knowledge of the materials to be excavated. Where subsurface conditions exhibit extreme variations over short distances, multiple borings at each shaft
location can reduce the risk of founding a shaft on soil instead of rock. For example, large-diameter, nonredundant
shafts in karstic limestone may require multiple borings at
each shaft location to determine that the entire base will be
founded in rock and to identify voids or zones of soil
beneath the base that may affect load-settlement behavior
of each shaft.
The draft 2006 Interim AASHTO LRFD Bridge Design
Specifications recommends the following for depth of borings
below anticipated tip elevations:

TABLE 2
RECOMMENDED FREQUENCY OF BORINGS,
DRILLED SHAFT FOUNDATIONS FOR
BRIDGES, UNCLASSIFIED EXCAVATION
Redundancy
Condition
Single-column, single
shaft foundations
Redundant, multipleshaft foundations
Redundant, multipleshaft foundations
Redundant, multipleshaft foundations

Shaft
Diameter
(m)
All
>1.8 m
(6 ft)
1.21.8 m
(46 ft)
<1.2 m
(4 ft)

Source: ONeill and Reese 1999.

Guideline
One boring per
shaft
One boring per
shaft
One boring per
two shafts
One boring per
four shafts

For shafts supported on or extending into rock, a minimum of


3 m of rock core, or a length of rock core equal to at least three
times the shaft diameter for isolated shafts or two times the maximum shaft group dimension, whichever is greater, shall be
extended below the anticipated shaft tip elevation to determine
the physical characteristics of rock within the zone of foundation
influence.

If the tip elevation changes at some point during the project,


additional drilling may be required to meet this recommendation. ONeill and Reese (1999) provide the following guidance
on boring depth. When the RQD is less than 50%, extend boring depths to at least 125% of the expected depths of the drilled
shaft bases plus two base diameters. If RQD values are greater
than approximately 50% at the planned base elevation, borings
only need be extended to the expected base elevation plus two
base diameters as long as the RQD remains above 50%. The
rationale is that it is not likely the shafts will need to be deepened once the actual strata are exposed. This approach requires
that foundation diameters and depths be estimated before the
boring program and that RQD be determined during drilling.
The approach described is only a general suggestion and local
geologic conditions may dictate other criteria for boring depths.
If in the course of design or construction it becomes necessary
to deepen the shafts, supplementary borings should be taken.
An available, but not widely used tool for subsurface investigation is to drill one or more large-diameter borings or to have
a drilled shaft contractor install a full-sized test excavation.
Large-diameter borings can be made with augers in soft rock
and with core barrels in hard rock. The sidewalls of the boring
or shaft can be examined directly (with appropriate safety measures) or with downhole cameras. Observations can then be
made of rock mass features, including degree of roughness and
general quality of the drilled surfaces, and fracture patterns.
Large-diameter holes provide access for obtaining high-quality
undisturbed samples and may be used for performing in situ
plate load tests to measure rock mass modulus. If a full-size
excavation is made by a drilled shaft contractor, information of
value to both engineers and contractors is obtained. In Figure 3, constructability is one of the items to be determined
during the site characterization. A full-sized excavation is the
most direct method for obtaining this information.
Downhole devices are available for borehole viewing and
photography, including borescopes, photographic cameras, and
television cameras. A visual image of rock in the sidewalls of
a boring provides information on structural features that may
add significantly to the overall picture of subsurface geology.
Advantages and disadvantages of some remote viewing
devices are discussed in Geotechnical Investigations (2001);
however, the technologies for borehole imaging are advancing
rapidly and the user should consult commercial providers for
the most up-to-date information. These devices are effective
for examining soft zones for which core may not have been
recovered, determination of dip and strike of important structural features, and viewing of cavities such as solution voids,
open joints, and lava tunnels in volcanic rocks.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

15

Borehole televiewers provide high-resolution images


showing rock mass structural and textural features and accurate measurement of dip and dip direction of structural
features without the use of oriented core. Optical televiewers (OTV) generate a high-resolution digital color image of
the inside of the borehole wall and are capable of resolving
fractures as narrow as 0.1 mm with a radial resolution of
1 degree (Eliassen et al. 2005). The OTV can be operated in
air- or fluid-filled boreholes; however, fluid requires thorough flushing before image acquisition is undertaken.
Acoustic televiewers (ATV) produce images of the borehole
wall based on the amplitude and travel time of acoustic signals
reflected from the borehole wall. A portion of the reflected
energy is lost in voids or fractures, producing dark bands on
the amplitude log. Travel time measurements allow reconstruction of the borehole shape, making it possible to generate
a 3-D representation of a borehole.
Both types of televiewers orient their image data using
a three-component fluxgate magnetometer and a threecomponent tilt meter incorporated into the tool. Before interpretation, the image is rotated to a common reference direction,
either magnetic north or the high side of the borehole. Planar
features that intersect the borehole wall produce sinusoidal
traces in the unwrapped, or 2-D, televiewer image. Using
the reference direction recorded during logging, sinusoids can
be analyzed to produce dip and dip directions of structural
features. Figure 9 shows OTV and ATV images of the same
borehole and illustrates some advantages of each device. The
OTV is able to provide a color image of the dike and excellent imaging of the texture of the granite. The ATV highlights
fracturing within the diorite. The California DOT (Caltrans)
reports using the ATV to provide very-high-resolution sonic
images in the format of a 3-D pseudo-core, as illustrated in
Figure 10.

FIGURE 9 Optical and acoustic televiewer images of a 50-cm


diorite dike in granite (Eliassen et al. 2005).

FIGURE 10 An acoustic television log (Caltrans 2005).

According to Eliassen et al. (2005), use of optical and


acoustic televiewer equipment is gaining popularity over
oriented coring techniques because it is generally less labor
intensive and is particularly useful where access or ability to
drill inclined holes is limited or where local drilling companies lack the equipment necessary to collect oriented cores.
However, to date, this technology is being applied to site characterization for rock slope engineering and underground
openings, and is not being used in foundation investigations.
Eliassen et al. (2005) note further that televiewer logs are best
used to supplement data obtained from quality rock coring,
which provides samples for laboratory testing, assessment of
joint and discontinuity planes, and correlation of lithologic
and geologic boundaries with geophysical data. The authors
suggest that drilling time and costs can be optimized with appropriate combinations of coring and less expensive air rotary
boreholes logged with OTV and ATV equipment. Borehole
televiewing may be most useful in rock-socket applications at
sites where the structural orientation of discontinuities is a
significant factor in foundation stability. For example, some
modes of bearing capacity failure (described in chapter three)
depend on the orientation of discontinuities in the rock mass
below the socket base. LaFronz et al. (2003) describe use of
OTV as part of the subsurface investigation for the Colorado
River Bridge at Hoover Dam. The primary purpose was to
obtain structural data to develop recommendations for excavation of cut slopes at the abutment foundations.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

16
TABLE 3
ROCK GROUPS AND TYPES

GEOLOGIC AND INDEX PROPERTIES OF ROCK

The most basic characterization of rock for engineering


purposes is a description of rock core based on visual and
physical examination. The International Society of Rock
Mechanics (ISRM) proposed a standardized method for
descriptions of rock masses from mapping and core logging
(Basic Geotechnical Description of Rock Masses 1981). A
summary of the ISRM method as given by Wyllie (1999) is
adopted in the FHWA manuals on subsurface investigations
and soil and rock properties (Mayne et al. 2001; Sabatini
et al. 2002) and is summarized here.
A rock mass is described in terms of five categories of
properties, as follows:
1. Rock Material Descriptiona. Rock type, b. Wall
strength, c. Weathering
2. Discontinuity Descriptiond. Type, e. Orientation, f.
Roughness, g. Aperture
3. Infillingh. Infilling type and width
4. Rock Mass Descriptioni. Spacing, j. Persistence, k.
Number of sets, l. Block size/shape
5. Groundwaterm. Seepage.
Each of the 13 parameters listed (a through m) is assigned
a description using standardized terminology. Descriptive
terms are given in Tables 3 through 6 and in Figure 11, which
is an example of a Key used for entering rock descriptions on
a coring log and includes details of several categories.
Rock Material Descriptors

Rock type is defined in terms of origin (igneous, sedimentary,


or metamorphic) and then further classified into one of the

Intrusive
(coarse-grained)
Granite
Syenite
Diorite
Diabase
Gabbro
Peridotite
Pegmatite
Clastic (sediment)
Shale
Mudstone
Claystone
Siltstone
Conglomerate
Limestone, oolitic

Igneous
Extrusive
(fine-grained)
Rhyolite
Trachyte
Andesite
Basalt

Sedimentary
(chemically formed)
Limestone
Dolomite
Gypsum
Halite

Diameter (mm)
>4.75
2.004.75
0.4252.00
0.0750.425
<0.075

Subangular

Subrounded
Rounded
Well-rounded

Nonfoliated
Quartzite
Amphibolite
Marble
Hornfels

rock types listed in Table 3 based on lithologic characteristics


that include color, fabric (microstructural and textural features), grain size and shape (Tables 4 and 5), and mineralogy.
Sedimentary rock descriptions should include bedding thickness (Table 6). The rock unit name, which may be a formal
name of a formation or an informal local name, should be
identified; for example, Bearpaw Shale or Sherman Granite.
Compressive strength of rock core can be evaluated using simple field tests with equipment commonly available
(knife, rock hammer, etc.) and summarized in the Key of
Figure 11 (Rock Strength) or evaluated from point load

Characteristic
Grain sizes are greater than popcorn kernels
Individual grains can be easily distinguished by eye
Individual grains can be distinguished by eye
Individual grains can be distinguished with difficulty
Individual grains cannot be distinguished by unaided eye

TABLE 5
TERMS TO DESCRIBE GRAIN SHAPE (for sedimentary rocks)
Description
Angular

(organic remains)
Chalk
Coquina
Lignite
Coal

Metamorphic
Foliated
Slate
Phyllite
Schist
Gneiss

TABLE 4
TERMS TO DESCRIBE GRAIN SIZE OF SEDIMENTARY ROCK
Description
Very coarse grained
Coarse grained
Medium grained
Fine grained
Very fine grained

Pyroclastic
Obsidian
Pumice
Tuff

Characteristic
Showing very little evidence of wear. Grain edges and corners are sharp. Secondary
corners are numerous and sharp.
Showing definite effects of wear. Grain edges and corners are slightly rounded off.
Secondary corners are slightly less numerous and slightly less sharp than in angular
grains.
Showing considerable wear. Grain edges and corners are rounded to smooth curves.
Secondary corners are reduced greatly in number and highly rounded.
Showing extreme wear. Grain edges and corners are smoothed off to broad curves.
Secondary corners are few in number and rounded.
Completely worn. Grain edges and corners are not present. No secondary edges or
corners are present.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

17
TABLE 6
TERMS TO DESCRIBE STRATUM
THICKNESS
Descriptive Term
Very thickly bedded
Thickly bedded
Thinly bedded
Very thinly bedded
Laminated
Thinly laminated

Stratum Thickness
>1 m
0.5 to 1.0 m
50 mm to 500 mm
10 mm to 50 mm
2.5 mm to 10 mm
<2.5 mm

tests or uniaxial compression tests conducted on specimens.


The rock strength descriptions given at the bottom of the
second page of the Key correspond to the seven categories
of rock strength, R0 through R6, of the ISRM (Basic Geotechnical Description of Rock Masses 1981), with R0 corresponding to extremely weak rock and R6 corresponding
to extremely strong rock. The degree of physical disintegration or chemical alteration of rock can be described by the
terms and abbreviations given in the Key. Weathering and
alteration reduces shear strength of both intact rock and
discontinuities.

FIGURE 11 Key for rock core description (sheet 1).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

18

FIGURE 11 (continued ) (sheet 2).

Discontinuity Descriptors

A discontinuity is defined as any surface across which any mechanical property of a rock mass is discontinuous. Discontinuity descriptors are summarized in Figure 11 (Key), items a
through g. Types of discontinuities include faults, joints, shear
planes, foliation, veins, and bedding. Orientation refers to the
measured dip and dip direction of the surface (or dip and
strike). Dip is defined as the maximum angle of the plane to
the horizontal and dip direction (strike) is the direction of the
horizontal trace of the line of dip measured clockwise from
north, in degrees. Determination of dip and dip direction from
core samples is possible using oriented coring techniques,
borehole televiewers, downhole cameras, or other devices
capable of establishing orientation of the discontinuity relative

to the core. Roughness and surface shape of joint surfaces is


best measured in the field on exposed surfaces at least 2 m in
length and can be described using the terms in the Key or
quantified in terms of a Joint Roughness Coefficient (Barton
1973). Aperture is the width of a discontinuity with no infilling and can be classified according to Box c of the Key.

Infilling

Infilling is the term for material separating adjacent rock


walls of discontinuities. Infilling is described in terms of its
type, amount, and width (Key). Additional laboratory testing
may be conducted to determine soil classification and shear
strength of infilling materials. Direct shear tests provide a

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

19

means to measure shear strength of joints with infilling, as


described by Wyllie and Norrish (1996). Infilling properties
vary widely and can have a significant influence on rock
mass strength (RMS), compressibility, and permeability.

L = 250 mm

RQD =

L=0
HIGHLY WEATHERED
DOES NOT MEET
SOUNDNESS REQUIREMENT

L=0
CENTER LINE
PIECES < 4"
& HIGHLY WEATHERED

Spacing is the perpendicular distance between adjacent discontinuities. Spacing has a major influence on seepage and
mechanical behavior and can be described using the terms in
Figure 11 (Key). Persistence refers to the continuous length
or area of a discontinuity and requires field exposures for its
determination.
The number of sets of intersecting discontinuities has a
major effect on RMS and compressibility. As the number of
sets increases, the extent to which the rock mass can deform
without failure of intact rock also increases. Field mapping
or observations made in exploratory pits or large excavations
provide the best opportunity to map multiple sets of discontinuities. Block size and shape is determined by spacing, persistence, and number of intersecting sets of discontinuities.
Descriptive terms include blocky, tabular, shattered, and
columnar, while size ranges from small (<0.0002 m3) to very
large (>8 m3).

Seepage

Field observations of seepage from discontinuities should


be described whenever it can be observed. The presence
and type of infilling controls joint permeability and should
be described wherever seepage is observed. Seepage can
range from dry to continuous flow under high pore water
pressure

Rock Quality Designation

A simple and widely used measure of rock mass quality is


provided by the RQD (rock quality designation, ASTM
D6032). RQD is equal to the sum of the lengths of sound
pieces of core recovered, greater than 100 mm (4 in.) in
length, expressed as a percentage of the length of the core
run. Originally introduced by Deere (1964), the RQD was
evaluated by Deere and Deere (1989), who recommended
modifications to the original procedure after evaluating its
field use. Figure 12 illustrates the recommended procedure.
Several factors must be evaluated properly for RQD to provide reliable results.
RQD was originally recommended for NX size core, but
can also be used with the somewhat smaller NQ wireline
sizes and with larger wire line sizes and other core sizes
up to 150 mm (6 in.). RQD based on the smaller BQ and
BX cores or with single-tube core barrels is discouraged
because of core breakage. Core segment lengths should be

L = 190 mm

L=0
< 4"
MECHANICAL
BREAK CAUSED
BY DRILLING
PROCESS

250 + 190 + 200


100%
1200

RQD = 53% (FAIR)

CORE RUN TOTAL LENGTH = 1 2 0 0 m m

Rock Mass Descriptors

RQD =

LENGTH OF SOUND
> 100 mm
PIECES
TOTAL CORE RUN LENGTH

CORE

L = 200 mm

L=0
NO RECOVERY

FIGURE 12 RQD determination of rock core (after Deere and


Deere 1989).

measured along the centerline or axis of the core, as shown


in Figure 12.
Only natural fractures such as joints or shear planes
should be considered when calculating RQD. Core breaks
caused by drilling or handling should be fitted together and
the pieces counted as intact lengths. Drilling breaks may be
identified by fresh surfaces. For some laminated rocks it may
be difficult to distinguish natural fractures from those caused
by drilling. For characterization of rock mass behavior relevant to foundation design it is conservative to not count the
length near horizontal breaks. RQD should be performed as
soon as possible after the core is retrieved to avoid the effects
of deterioration, which may include slaking and separation of
core along bedding planes, especially in moisture-sensitive
rocks like some shales. It is also desirable because RQD is a
quantitative measure of core quality at the time of drilling
when the rock core is fresh and most representative of in
situ conditions.
Rock assigned a weathering classification of highly weathered or above should not be included in the determination of
RQD. RQD measurements assume that core recovery is at or
near 100%. As core recovery varies from 100%, explanatory
notes may be required to describe the reason for the variation
and the effect on RQD. In some cases, RQD will have to be
determined on the basis of total length of core recovered, rather
than on the length of rock cored. One state (Florida) uses percent core recovery as an index of rock quality in limestone.
A general description of rock mass quality based on RQD
is given here. Its wide use and ease of measurement make it
an important piece of information to be gathered on all core
holes. Taken alone, RQD should be considered only as an

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

20

approximate measure of overall rock quality. RQD is most


useful when combined with other parameters accounting for
rock strength, deformability, and discontinuity characteristics. As discussed in subsequent sections of this report, many
of the rock mass classification systems in use today incorporate RQD as a key parameter.
Rock Mass Description
Excellent
Good
Fair
Poor
Very Poor

RQD
90100
7590
5075
2550
<25

ENGINEERING PROPERTIES OF ROCK


Laboratory Tests on Intact Rock

Intact rock refers to the consolidated and cemented assemblage of mineral particles forming the rock material, excluding the effects of macro-scale discontinuities such as
joints, bedding planes, minor faults, or other recurrent planar fractures. The term rock mass is used to describe the system comprised of intact rock and discontinuities. The characteristics of intact rock are determined from hand
specimens or rock core. Properties of intact rock required for
proper characterization of the rock mass and that are relevant to foundation design include strength and deformability. For some rock types, the potential for degradation on exposure to atmospheric conditions may also need to be
evaluated. Some design methods incorporate properties of
intact rock directly; for example, correlations between ultimate unit side resistance and uniaxial compressive strength.
However, most analytical treatments of foundation capacity
and load-deformation response incorporate the strength and
deformability of intact rock into rock mass models that also

account for the effects of discontinuities, rock quality, and


other factors.
Table 7 lists the laboratory tests for intact rock most commonly done for foundation design and gives the ASTM
Standard Designation for each test. More thorough coverage
of laboratory testing of intact rock is given by Mayne et al.
(2001), the Rock Testing Handbook (1993), and the
AASHTO Manual on Subsurface Investigations (1988).
Engineering properties of intact rock that are used most
often for foundation design are uniaxial compressive
strength (qu) and elastic modulus (ER). The compressive
strength of intact rock is determined by applying a vertical
compressive force to an unconfined cylindrical specimen
prepared from rock core. The peak load is divided by the
cross-sectional area of the specimen to obtain the uniaxial
compressive strength (qu). The ASTM procedure (D2938)
specifies tolerances on smoothness over the specimen
length, flatness of the ends, the degree to which specimen
ends are perpendicular to the length, and length-to-diameter
ratio. Uniaxial compressive strength of intact rock is
used in empirical correlations to evaluate ultimate side
and base resistances under axial loading; ultimate limit
pressure under lateral loading; and, by contractors, to assess constructability.
Elastic modulus of intact rock is measured during conduct
of the uniaxial compression test by measuring deformation as
a function of load. It is common to measure both axial and diametral strain during compression to determine elastic modulus and Poissons ratio. Test procedures are given in ASTM
Standard (D3148) and discussed further by Wyllie (1999). It
is important to note that the ASTM procedure defines several
methods of determination of modulus, including tangent
modulus at a specified stress level, average modulus over the

TABLE 7
COMMON LABORATORY TESTS FOR INTACT ROCK
Test
Category
Uniaxial
compression

Name of Test and ASTM Designation


Unconfined compressive strength of intact
rock core specimen (D2938)

Split tensile

Splitting tensile strength of intact rock core


specimens (D3967)

Point load
strength

Determination of the point load strength index


of rock (D5731)

Direct shear

Laboratory direct shear strength tests for rock


specimens under constant normal stress
(D5607)
Elastic moduli of intact rock core specimens in
uniaxial compression (D3148)

Strengthdeformation
Durability

Slake durability of shales and similar weak


rocks (D4644)

Comments
Primary test for strength and deformability
of intact rock; input parameter for rock mass
classification systems
Splitting tensile strength of a rock disk under
a compression line load
Index test for rock strength classification;
can be performed in field on core pieces
unsuitable for lab testing
Applies to intact rock strength or to shear
strength along planes of discontinuities,
including rockconcrete interface
Youngs modulus from axial stressstrain
curve; Poissons ratio can also be
determined
Index test to quantify the durability of weak
rocks under wetting and drying cycles with
abrasion

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

21

linear portion of the stressstrain curve, and secant modulus


at a fixed percentage of maximum strength. For rocks that
exhibit nonlinear stressstrain behavior, these methods may
provide significantly different values of modulus and it is
important to note which method was used when reporting values of modulus.
The point load test is conducted by compressing a core
sample or irregular piece of rock between hardened steel
cones (Figure 13), causing failure by the development of tensile cracks parallel to the axis of loading. The uncorrected
point load strength index is given by
Is = P/D2

(4)

where P = load at rupture, and D is the distance between the


point loads. The point load index is reported as the point load
strength of a 50 mm core. For other specimen sizes a correction factor is applied to determine the equivalent strength of
a 50 mm specimen. The point load index is correlated to uniaxial compressive strength by
qu = C Is(50)

(5)

where qu is the unconfined compressive strength, Is(50) is the


point load strength corrected to a diameter of 50 mm, and

FIGURE 13 Point load test setup.

C is a correlation factor that should be established on a sitespecific basis by conducting a limited number of uniaxial
compression tests on prepared core samples. If a site-specific
value of C is not available, the ASTM Standard recommends
approximate values based on core diameter. For a 54 mm
core (NX core size), the recommended value of C is 24. The
principal advantages of the point load test are that it can
be carried out quickly and inexpensively in the field at the
site of drilling and that tests can be conducted on irregular
specimens without the preparation required for uniaxial compression tests.
Split tensile strength (qt) of rock (ASTM D4644) is determined by compressing a cylindrical disk under a compressive
line load. Split tensile strength has been correlated with unit
side resistance; for example, by McVay et al. (1992) for
drilled shafts in Florida limestone.
Direct shear testing is applicable to determination of the
MohrCoulomb shear strength parameters cohesion, c, and
friction angle, , of discontinuity surfaces in rock (ASTM
D5607). Shear strength of discontinuities may govern capacity in certain conditions; for example, base capacity of socketed foundations when one or two intersecting joint sets are
oriented at an intermediate angle to horizontal. The other notable application of this test is in simulating the shear
strength at the rockconcrete interface for evaluation of side
resistance of socketed shafts under axial loading. However,
for this application, the constant normal stiffness (CNS) direct shear test described by Johnston et al. (1987) is more applicable. Instead of a constant normal load, normal force is
applied through a spring that increases or decreases the applied force in proportion to the magnitude of normal displacement (dilation). Dilatancy of the interface is a major
factor controlling strength and stiffness of socketed shafts
under axial load.
The slake durability test (ASTM D4644) provides an index
for identifying rocks that will weather and degrade rapidly.
The test is appropriate for argillaceous sedimentary rocks
(mudstone, shale, clayshales) or any weak rock. Representative rock fragments are placed in a wire mesh drum and dried
in an oven to constant weight. The drum is partially submerged in water and rotated at 20 revolutions per minute for
a period of 10 min. The drum and its contents are then dried a
second time and the loss of weight is recorded. The test cycle
is repeated a second time and the slake durability index, ID, is
calculated as the ratio (reported as a percentage) of final to initial dry weights of the sample. Rocks with ID < 60 are considered prone to rapid degradation and may indicate a susceptibility to degradation of the borehole wall when water is
introduced during drilling, potentially leading to formation of
a smear zone. Hassan and ONeill (1997) define the smear
zone as a layer of soil-like material along the socket wall and
demonstrate that smearing can have a significantly negative
effect on side load transfer of shafts in argillaceous rock.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

22
In Situ Tests for Rock

In situ testing can be used to evaluate rock mass deformation


modulus and, in some instances, RMS. In situ testing methods with potential applications to rock-socket design are
presented in Table 8. In situ testing of rock is not performed
routinely for rock-socket design by most of the agencies
surveyed for this study. The survey responses indicate that
five state DOTs currently use the pressuremeter test (PMT)
to obtain design parameters. Of these, all five use the test
to obtain rock mass modulus. One state reported the use of
PMT to evaluate RMS in weak rocks. Four states use the
PMT for correlating test results with the parameters that define p-y curves for analysis of shafts under lateral loading
(chapter four). The term dilatometer is also used to describe a
pressuremeter intended for use in rock but should not be confused with the flat plate dilatometer used for in situ testing of
soil. One state (Massachusetts) reported using the borehole
jack to measure rock mass modulus. No states reported using
the plate load test for rock-socket design. Information on
conduct and interpretation of the tests identified in Table 8
and other in situ tests for rock are given in the relevant
ASTM standards, Rock Testing Handbook (1993) and Mayne
et al. (2001).
Heuze (1980) investigated the effect of test scale on the
modulus of rock masses. Several types of field tests, including borehole jack and plate load tests at different scales, were
included and results were compared with those of laboratory
compression tests. It was observed that in situ rock mass
modulus values generally range from 20% to 60% of intact

rock modulus from laboratory uniaxial compression tests.


The borehole jack was recommended as a field test that, with
proper analysis (Heuze 1984), yields values of rock mass
modulus that are consistent with results from large plate bearing tests. The borehole jack designed for NX sized borings
(75 mm or 3 in. diameter) affects a test volume of approximately 0.14 m3 (5 ft3). Borehole jack devices are available
commercially with limit pressures of up to 69 MPa, allowing
the test to reach stress levels beyond the elastic limit and, for
some weak rock masses, to ultimate strength.
Studies on the use of PMTs for determination of rock mass
modulus include those of Rocha et al. (1970), Bukovansky
(1970), Georgiadis and Michalopoulos (1986), and Littlechild
et al. (2000). Results have been mixed, with some researchers indicating a high degree of agreement between PMT
modulus and other in situ tests (e.g., Rocha et al. 1970) and
others reporting PMT modulus values significantly lower
than modulus measured by plate-load and borehole jack tests
(e.g., Bukovansky 1970). Littlechild et al. (2000) concluded
that PMTs, using the Cambridge High Pressure Dilatometer,
were not useful for determination of rock mass modulus for
design of deep foundations in several rock types in Hong
Kong. In strong and massive rocks such as metasiltstone and
tuff, the device did not have sufficient capacity to measure
modulus, which typically was around 10 GPa. In highly fractured granodiorite, membrane failures were problematic.
Commercially available pressuremeter devices for rock are
currently limited to maximum pressures of around 30 MPa.
Additional discussion of rock mass modulus is presented
later in this chapter.

TABLE 8
IN SITU TESTS WITH APPLICATIONS TO ROCK-SOCKET DESIGN
Method
Pressuremeter
(includes devices
referred to as
rock dilatometer)

Procedure
Pressuremeter is lowered to the
test elevation in a prebored
hole; flexible membrane of
probe is expanded exerting a
uniform pressure on the
sidewalls of the borehole
Jacks exert a unidirectional
pressure to the walls of a
borehole by means of two
opposed curved steel platens

Rock Properties
Rock mass modulus;
rock mass strength
in weak rocks
ASTM D4719

Limitations/Remarks
Test affects a small area of rock
mass; depending on joint
spacing, may or may not
represent mass behavior; limited
to soft or weak rocks

Rock mass modulus;


rock mass strength
in weak rocks
ASTM D4971

Plate load test

Load is applied to a steel plate


or concrete foundation using a
system of hydraulic jacks and a
reaction frame anchored to the
foundation rock

Rock mass modulus;


rock mass strength
in weak rocks

Texas cone
penetration test

Steel cone is driven by a drop


hammer; number of blows per
300 mm of penetration is TCPT
N-value; depth of penetration
per 100 blows is penetration
resistance (PR)

Correlated to
compressive strength
of weak rocks
encountered in
Texas and Oklahoma

Measured modulus value must


be corrected to account for
stiffness of steel platens; test
method can be used to provide
an estimate of anisotropy
Loaded area is limited, so may
not be effectively testing rock
mass if joints are widely spaced;
modulus values corrected for
plate geometry, effect of rock
breakage, rock anisotropy, and
steel plate modulus; not common
for deep foundations
Limitations similar to those of
Standard Penetration Test;
currently used by Texas and
Oklahoma DOTs for direct
correlation to side and base
resistance of shafts in weak rock

Borehole jack

Notes: Adapted from Geotechnical Engineering Circular No. 5 (Sabatini et al. 2002).
TCPT = Texas Cone Penetration Test.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

23

An example of an in situ test that is used in a specific region of the country is the Texas Cone Penetration Test
(TCPT). A 76-mm-diameter solid steel cone is driven by a
77 kg (170 lb) drop hammer. The number of blows required
to drive 300 mm (12 in.) is recorded and the results are given
in one of two ways: (1) number of blows per 300 mm of penetration or TCPT N-value, or (2) the depth of penetration per
100 blows, referred to as the penetration resistance or PR.
The Texas and Oklahoma DOTs use empirical correlations
between the TCPT parameters and drilled shaft side and base
resistances in soil and soft rock. The test procedure and
correlations are available in the Texas DOT Geotechnical
Manual, which can be accessed online. Some researchers
have developed empirical correlations between TCPT measurements and properties of soft rock. For example, Cavusoglu et al. (2004) show correlations between compressive
strength of upper Cretaceous formation clay shales (UU triaxial tests) and limestone (unconfined compression) and PR
measurements conducted for Texas DOT projects. The correlations are highly formation-dependent and exhibit a high
degree of scatter, but provide first order estimates of rock
strength based on TCPT resistance in formations where sample recovery is otherwise difficult.
In addition to the tests identified as being applicable to
rock, it is common practice to use in situ tests for soil to
define the contact boundary between soil and rock. Of the
agencies surveyed, 21 reported using the Standard Penetration Test (SPT) and 3 reported using the Cone Penetration
Test (CPT) to define the top-of-rock elevation. Refusal of
the SPT or CPT penetration is the method most often used to
identify rock. Limitations of this approach include the possibility of mistaking cobbles or boulders for the top-of-rock
and the lack of consistency in SPT blowcounts in weak or
weathered rock.
Six states reported using the SPT in soft or weak rock to
obtain rock properties (unconfined compressive strength) or
for correlating SPT N-values directly to design parameters,
principally unit side resistance. For example, the Colorado
SPT-Based Method is used by the Colorado DOT to establish design values of both unit side resistance and base resistance for shafts socketed into claystones when the material
cannot be sampled in a way that provides intact core specimens adequate for laboratory uniaxial compression tests
(Abu-Hejleh et al. 2003). ONeill and Reese (1999) correlate
unit side resistance with N-values for shafts in cohesionless
IGMs, defined as materials with N > 50. Direct correlations
between design parameters and N values are considered further in chapter three.

Rock Mass Classification

Several empirical classification systems have been proposed


for the purpose of rating rock mass behavior. The most
widely used systems are the Geomechanics Classification

described by Bieniawski (1976, 1989) and the Rock Quality


Tunneling Index described by Barton et al. (1974). Both systems were developed primarily for application to tunneling
in rock, but have been extended to other rock engineering
problems. The application of classification systems to rocksocket design has been limited to correlations between classification parameters and RMS and deformation properties.
To facilitate such correlations, Hoek et al. (1995) introduced
the GSI. Relationships were developed between GSI and the
rock mass classifications of Bieniawski and Barton et al. The
principal characteristics of the two classification systems are
summarized, followed by a description of their relationship
to GSI. For more detailed discussion, including limitations
and recommended applications, consult the original references and Hoek et al. (1995, 2002).
The Geomechanics Classification is based on determination of the RMR, a numerical index determined by summing
the individual numerical ratings for the following five categories of rock mass parameters:

Strength of intact rock,


Drill core quality (in terms of RQD),
Spacing of discontinuities,
Condition of discontinuities, and
Groundwater conditions.

An adjustment is made to the RMR for the degree to


which joint orientation may be unfavorable for the problem
under consideration. The classification system is presented in
Table 9. Based on the RMR value, a rock mass is identified
by one of five rock mass classes, ranging from very poor rock
to very good rock. The draft 2006 Interim AASHTO LRFD
Bridge Design Specifications recommends determination of
RMR for classification of rock mass in foundation investigations. Seventeen states reported using RMR either always or
sometimes for rock mass classification associated with
drilled shaft design.
Barton and co-workers at the Norwegian Geotechnical Institute proposed a Tunneling Quality Index (Q) for describing
rock mass characteristics and tunnel support requirements
(Barton et al. 1974). The system is commonly referred to as
the NGI-Q system or simply the Q-system. The numerical
value of the index Q varies on a log scale from 0.001 to 1,000
and is defined as:
Q=

RQD J r
J

w
Jn
J a SRF

where
RQD = rock quality designation,
Jn = joint set number,
Jr = joint roughness number,
Ja = joint alteration number,
Jw = joint water reduction factor, and
SRF = stress reduction factor.

Copyright National Academy of Sciences. All rights reserved.

(6)

Rock-Socketed Shafts for Highway Structure Foundations

24
TABLE 9
GEOMECHANICS CLASSIFICATION SYSTEM FOR DETERMINATION OF ROCK MASS RATING (RMR)

2
3
4

A. Classification Parameters and Their Ratings (after Bieniawski 1989)


Ranges of Values
Parameter
Strength
of
Point load
strength
intact
>10
410
24
12
index, MPa
rock
material
Uniaxial
comp.
>250
100250
50100
2550
strength,
MPa
15
12
7
4
Rating
90100
7590
5075
2550
Drill core quality, RQD (%)
20
17
13
8
Rating
>2 m
0.62 m
200600 mm
60200 mm
Spacing of discontinuities
20
15
10
8
Rating
Condition of discontinuities
Slightly
Slickensided
rough
Very rough
Slightly rough
surfaces or
surfaces,
surfaces,
surfaces,
gouge <5 mm
not continuous,
separation
separation <1
thick or
<1 mm,
no separation,
mm, highly
joints open 1 to
unweathered
slightly
weathered
5 mm
wall rock
weathered
walls
continuous
walls
30
25
20
10
Rating
GroundInflow per 10 m
None
<10
1025
25125
water
tunnel length
Ratio: Joint
water pressure/
0
<0.1
0.10.2
0.20.5
major principal

For this low


range, uniaxial
comp. test is
preferred
525
2

15

1
<25
3
<60 mm
5

<1
0

Soft gouge >5


mm thick or
separation >5
mm
continuous
0
>125

>0.5

stress
General
conditions
Rating
Strike and dip
orientations
Ratings
RMR
Class Number
Description

Foundations

Completely dry

Damp

15
10
7
B. Rating Adjustment for Joint Orientations
Very
Favorable
Fair
favorable
2
7
0
C. Rock Mass Classes Determined from Total Ratings
100 to 81
80 to 61
60 to 41
I
II
III
Very good
rock
Good rock
Fair rock

Three states reported using the Q-system in connection


with rock-socket design. A modified Tunneling Quality Index (Q') is utilized to determine the GSI, as described
subsequently.
The Geomechanics Classification can be used to estimate
the value of GSI for cases where RMR is greater than 23, as
follows:
GSI = RMR89 5

(7)

in which RMR89 is the RMR according to Bieniawski (1989)


as presented in Table 9. For RMR89 values less than 23, the
modified (Q) is used to estimate the value of GSI, where:
Q' =

RQD Jr

Jn
Ja

Wet

(8)

Dripping

Flowing

Unfavorable
15

Very
Unfavorable
25

40 to 21
IV

<20
V

Poor rock

Very poor rock

GSI = 9LogeQ' + 44

(9)

Table 10 gives the values of the parameters used to evaluate


Q' by Eq. 8.
Engineering Properties of Rock Mass

Shear Strength
Geotechnical evaluation of foundation ultimate capacity under axial and lateral loading is calculated on the basis of shear
strength along assumed failure surfaces in the rock or at the
concreterock interface. Depending on the failure mode, the
strength may need to be defined at one of three levels: (1) intact rock, (2) along a discontinuity, and (3) representative of
a highly fractured rock mass. Figure 14 illustrates these cases
for a socketed foundation in rock. For example, bearing

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

25
TABLE 10
JOINT PARAMETERS USED TO DETERMINE Q'
1. No. of Sets of Discontinuities = Jn
Massive
0.5
One set
2
Two sets
4
Three sets
9
Four or more sets
15
Crushed rock
20

3. Discontinuity Condition & Infilling = Ja


3.1 Unfilled cases
Healed
0.75
Stained, no alteration
1
Silty or sandy coating
3
Clay coating
4
3.2 Filled discontinuities
Sand or crushed rock infill
4
Stiff clay infilling <5 mm
6
Soft clay infill <5 mm thick
8
Swelling clay <5 mm
12
Stiff clay infill >5 mm thick 10
Soft clay infill >5 mm thick
15
Swelling clay >5 mm
20

2. Roughness of Discontinuities = Jr
Noncontinuous joints
4
Rough, wavy
3
Smooth, wavy
2
Rough, planar
1.5
Smooth, planar
1
Slick and planar
0.5
Filled discontinuities
1

*Note: Add +1 if mean joint spacing > 3 m. Modified from Barton et al. (1974).

capacity at the base of a socketed foundation in massive rock


would be evaluated in terms of the strength of the intact rock.
If the rock has regular discontinuities oriented as shown in
level 2, base capacity may be controlled by the strength along
the joint surfaces. If the rock is highly fractured (level 3),
bearing capacity would have to account for the overall
strength of the fractured mass.

most common test for intact rock is the uniaxial (unconfined)


compression test, which can be considered a special case of
triaxial testing with zero confining stress. The strength parameter obtained is the uniaxial compressive strength, qu,
which is related to the MohrCoulomb strength parameters by

For each of the three cases, shear strength may be expressed within the framework of the MohrCoulomb failure
criterion, where shear strength () is given by

However, the strength of intact rock is normally given simply


in terms of qu. Stability analyses of rock sockets governed by
massive rock are normally evaluated directly in terms of qu.
When rock core is not sufficient for uniaxial compression
testing, or sometimes for convenience, qu is correlated to results of point load tests. Uniaxial compressive strength is also
one of the parameters used for evaluating the strength of
highly fractured rock masses, as discussed later.

= c' + ' tan '

(10)

in which c' = effective stress cohesion intercept, ' = effective stress angle of friction, and ' = effective normal stress
on the failure plane. Evaluation of shear strength for each of
the three cases is summarized as follows.
For intact rock the parameters c' and ' can be determined
from laboratory triaxial shear tests on specimens prepared
from core samples. Triaxial testing procedures are given by
ASTM D2664 and AASHTO T226. The survey of state
DOTs indicates that triaxial testing is not used routinely. The

qu = 2c tan (45 + 12 )

(11)

Shear strength of discontinuities can be determined using


laboratory direct shear tests. The apparatus is set up so that
the discontinuity surface lies in the plane of shearing between
the two halves of the split box. Both peak and residual values of the strength parameters (c' and ') are determined.
Discussion of direct shear testing of discontinuities, including its limitations, is given by Wyllie and Norrish (1996).
For a planar, clean fracture (no infilling), the cohesion is
zero and the shear strength is defined only by the friction angle.
The roughness of the surface has a significant effect on the
value of friction angle. If the discontinuity contains infilling,
the strength parameters will be controlled by the thickness and
properties of the infilling material. Compilations of typical
representative ranges of strength parameter values for discontinuities are summarized in Mayne et al. (2001). The survey results indicate that direct shear testing of joints is not conducted
routinely by DOT agencies for rock-socket design.

Shear failure
along joint

(a) Massive rock

(b) Jointed rock

(c) Highly fractured rock

FIGURE 14 Base failure modes illustrating different operational


shear strength conditions.

For intact rock masses and for fractured or jointed rock


masses, Hoek and Brown (1980) proposed an empirical criterion for characterizing RMS. Since its appearance, this criterion
has been applied widely in practice and considerable experience

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

26

has been gained for a range of rock engineering problems. Based


on these experiences, the criterion has undergone several stages
of modification, most significantly by Hoek and Brown (1988),
Hoek et al. (1995, 2002), and Marinos and Hoek et al. (2000).
The nonlinear RMS is given by:
'

1' = 3' + qu mb 3 + s

qu

(12)

where
'1 and '3 = major and minor principal effective stresses,
respectively;
qu = uniaxial compressive strength of intact rock;
and
mb, s, and a are empirically determined strength parameters for the rock mass.
The value of the constant m for intact rock is denoted by
mi and can be estimated from Table 11. Hoek and Brown

(1988) suggested that the constants mb, s, and a could be related


empirically to the RMR described previously. Hoek et al. (1995)
noted that this process worked well for rock masses with RMR
greater than about 25, but not well for very poor rock masses.
To overcome this limitation, the GSI was introduced. Suggested relationships between GSI and the parameters mb/mi, s,
and a, according to Hoek et al. (2002) are as follows:
GSI 100
mb
= exp
28 14 D
mi

(13)

GSI 100
s = exp
9 3D

(14)

a=

GSI
20
1 1 15

+ e
e 3

2 6

(15)

in which D is a factor that depends on the degree of disturbance


to the rock mass caused by blast damage and stress relaxation.

TABLE 11
VALUES OF THE CONSTANT mi BY ROCK GROUP (Hoek et al. 1995)
Rock
Type

Class

Group

Clastic

Texture
Coarse

Medium

Fine

Very fine

Conglomerate
(22)

Sandstone
19

Siltstone
9

Claystone
4

Sedimentary

<------------

Graywacke -------------->
(18)
<--------------- Chalk ----------------->
7

Organic

<----------------- Coal ----------------->


(821)

Non-clastic
Carbonate

Breccia
(20)

Metamorphic

Chemical

Micritic
limestone
8

Gypstone
16

Anhydrite
13

Non-foliated

Marble
9

Hornfels
(19)

Quartzite
24

Slightly foliated

Migmatite
(30)

Amphibolite
31

Mylonites
(6)

Foliated*

Gneiss
33

Schists
(10)

Phyllites
(10)

Slate
9

Granite
33

Rhyolite
(16)

Obsidian
(19)

Granodiorite
(30)

Dacite
(17)

Diorite
(28)

An desite
19

Light

Igneous

Sparitic
limestone
(10)

Dark

Gabbro
27

Dolerite
(19)

Basalt
(17)

Breccia
(18)

Tuff
(15)

Norite
22
Extrusive pyroclastic type

Agglomerate
(20)

*These values are for intact rock specimens tested normal to foliation. The value of mi will be significantly different if failure occurs
along a foliation plane.
Note: Values in parentheses are estimates.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

27

The damage factor D ranges from zero for undisturbed in situ


rock masses to 1.0 for very disturbed rock masses. Hoek et al.
(2002) provide guidance on values of D for application to tunnel and rock slope problems, but no work has been published
relating D to drilled shaft construction.
Some problems involving fractured rock masses (e.g., bearing capacity) are more readily analyzed in terms of the
MohrCoulomb strength parameters than in terms of the
HoekBrown criterion. Hoek and Brown (1997) noted that
there is no direct correlation between the two sets of strength
parameters. However, they describe a procedure that involves
simulating a set of triaxial strength tests using the HoekBrown
criterion (Eq. 12) then fitting the MohrCoulomb failure envelope to the resulting Mohrs circles by regression analysis.
Values of the strength parameters c' and ' defining the intercept and tangent slope of the envelope (which is nonlinear) can thus be determined. Hoek et al. (2002) presented the
following equations for the angle of friction and cohesive
strength of fractured rock masses:

6amb ( s + mb '3n )
' = sin 1
a 1
(
(
)
)
2 1 + a 2 + a + 6amb ( s + mb '3n )
a 1

c' =

qu [(1 + 2a ) s + (1 a ) mb '3n ]( s + mb '3n )

(1 + a ) ( 2 + a ) 1 + 6amb ( s + mb '3n )
(1 + a ) ( 2 + a )

a 1

(16)

a 1

(17)

Applications of the HoekBrown criterion to rock-socket


design are discussed further in chapter three (bearing
capacity) and chapter four (lateral capacity). The draft 2006
Interim AASHTO LRFD Bridge Design Specifications recommend the HoekBrown strength criterion for RMS characterization, but the earlier version (Hoek and Brown 1988)
is presented rather than the updated version based on GSI.
Deformation Properties
Rock mass deformation properties are used in analytical
methods for predicting the load-deformation behavior of
rock-socketed foundations under axial and lateral loads. The
parameters required by most methods include the modulus of
deformation of the rock mass, EM, and Poissons ratio, v.
Methods for establishing design values of EM include:
Estimates based on previous experience in similar rocks
or back-calculated from load tests,
Correlations with seismic wave velocity propagation
(e.g., Eqs. 13),
In situ testing, and
Empirical correlations that relate EM to strength or modulus values of intact rock (qu or ER) and/or rock mass
characteristics.
Compilations of typical values of rock mass modulus
and Poissons ratio are given in several sources, including

Kulhawy (1978), Wyllie (1999), and the AASHTO LRFD


Bridge Design Specifications (2004). These values should
be considered as general guidelines to expected ranges of
values for different rock types and serve to illustrate the
magnitude of variation that is possible. Rock mass modulus
can vary from less than 1 MPa to greater than 100 GPa and
depends on intact rock modulus, degree of weathering, and
characteristics of discontinuities. Compiled values provide
guidance for very preliminary evaluations, but should not be
relied on for final design. Values of Poissons ratio exhibit a
narrow range of values, typically between 0.15 and 0.3.
Various authors have proposed empirical correlations
between rock mass modulus and other rock mass properties. Table 12 presents, in chronological order, some of the
most widely cited expressions found in the literature. The
earliest published correlations (expressions 1 and 2 of
Table 12) relate EM to modulus of intact rock, ER, and RQD.
In subsequent correlations (expression 3), RQD is replaced
by RMR, providing a more comprehensive empirical approach because six rock mass parameters (including RQD)
are incorporated to evaluate the RMR. This was followed
by correlations relating EM directly to rock mass indexes,
including RMR and Q (expressions 4, 5, and 6). Hoek et al.
(1995) show the graph given in Figure 15 with curves given
by expressions 4, 5, and 6 of Table 12, along with case history observations. The figure suggests that expression 4 of
Table 12 provides a reasonable fit to the available data and
offers the advantage of covering a wider range of RMR values than the other equations. The draft 2006 Interim
AASHTO LRFD Bridge Design Specifications recommend
use of either expression 4 of Table 12 or a method recommended by ONeill et al. (1996) based on applying a modulus reduction ratio (EM/ER) given as a function of RQD in
Table 13.
Beginning with Hoek and Brown (1997), proposed correlation equations have been based on relating EM to GSI and
properties of intact rock, either uniaxial compressive strength
(qu) or intact modulus (ER). In expression 7, EM is reduced
progressively as the value of qu falls below 100 MPa. This reduction is based on the reasoning that deformation of better
quality rock masses is controlled by discontinuities, whereas
for poorer quality rock masses deformation of the intact
rock pieces contributes to the overall deformation process
(Hoek and Brown 1997). The version given in Table 12 is
updated by Hoek et al. (2002) to incorporate the damage
factor, D.
The final correlation (expression 8) in Table 12 was proposed based on analyses by Yang (2006). Figure 16 shows a
comparison of the regression equation (expression 8) to data
from field observations of Bieniawski (1978) and Serafim and
Pereira (1983), as well as modulus values measured by PMTs
reported by Yang (2006). Expression 8 was applied to derivation of p-y curves for analysis of laterally loaded rock
sockets, described further in chapter four. Additional discussion of empirical equations for rock mass modulus and their

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

28
TABLE 12
EMPIRICAL METHODS FOR ESTIMATING ROCK MASS MODULUS
Expression

Notes/Remarks

Reference

1. EM = ER[0.0231(RQD) 1.32]

Reduction factor on intact


rock modulus;
EM/ER > 0.15

Coon and Merritt


(1969); LRFD Bridge
Design . . . (2004)

2. For RQD < 70: EM = ER (RQD/350)


For RQD > 70: EM = ER [0.2 + (RQD 70)/37.5]

Reduction factor on intact


rock modulus

Bieniawski (1978)

Reduction factor on intact


rock modulus;
EM/ER < 1.0

Kulhawy (1978)

3. E = E 0.1 +
M
R

RMR
1150 11 .4RMR

5. EM (GPa) = 2 RMR 100

45 < RMR < 90

Serafim and Pereira


(1983)
Bieniawski (1984)

6. EM (GPa) = 25 log10 Q

1 < Q < 400

Hoek et al. (1995)

Adjustment to Serafim
and Pereira to account for
rocks with qu < 100 MPa;
note qu in MPa

Hoek and Brown


(1997); Hoek et al.
(2002)

Reduction factor on intact


modulus, based on GSI

Liang and Yang (2006)

4. E M (GPa ) = 10

RMR 10
40

7. E (GPa ) = 1 D
M
2

E M (GPa ) = 1

0 < RMR < 90

qu
10
100

D
10
2

GSI 10
40

GSI 10
40

for qu < 100 MPa


for qu > 100 MPa

GSI
8. E = E R e 21.7
M

100

Notes: ER = intact rock modulus, EM = equivalent rock mass modulus, RQD = rock quality designation, RMR =
rock mass rating, Q = NGI rating of rock mass, GSI = geological strength index, qu = uniaxial compressive strength.

application to foundation engineering is given by Littlechild


et al. (2000), Gokceoglu et al. (2003), and Yang (2006).
Rock mass modulus is a key parameter for rock-socket
load-deformation analysis, which is a key step in the design
process depicted in Figure 3. Several methods are identified in
this chapter for establishing values of EM. These include geophysical methods based on p-wave and s-wave velocities (Eqs.
1 and 2) or shear wave frequency (Eq. 3), in situ testing methods (Table 8), and the correlation equations given in Table 12.
The survey shows that correlation equations are the most
widely used method for estimating modulus for rock-socket
design, followed by in situ testing. The most common in situ test
(used by five states) is pressuremeter (rock dilatometer), with
a single state (Massachusetts) reporting use of the borehole

jack test. At least three other states using PMT for rock did
not respond to the survey. The principal limitation of in situ
testing is whether the volume of rock being tested is representative of the in situ rock mass. Factors such as degree of rock
disturbance, anisotropy, and spacing of discontinuities relative
to the dimensions of the apparatus will determine the degree
to which test results represent the response of rock mass to
foundation loading. As noted earlier in this chapter, rock mass
modulus measured by pressuremeter shows varying levels of
agreement with other in situ testing methods. The full range of
application and limitations of PMTs for rock mass modulus
and its application to rock-socket design have yet to be determined. Correlation equations for rock mass modulus have
evolved over the years as illustrated by the relationships summarized in Table 12. Correlations are attractive because they
are based on more easily measured properties of intact rock
and rock mass indexes, but caution must be exercised because
most of the correlations were developed specifically for applications to tunneling. Calibration studies aimed at the application of correlation equations for rock mass modulus to loaddeformation analysis of rock-socketed foundations are largely
lacking at the present time. Studies by Littlechild et al. (2000)
and Liang and Yang (2006) are exceptions and illustrate the
type of additional work that is needed.
TABLE 13
ESTIMATION OF MODULUS RATIO (EM /ER)
BASED ON RQD (ONeill et al. 1996)

FIGURE 15 Rock mass modulus versus rock mass rating


(Hoek et al. 1995).

RQD (percent)
100
70
50
20

EM/ER
Closed Joints
Open Joints
1.00
0.60
0.70
0.10
0.15
0.10
0.05
0.05

RQD = rock quality designation.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

29

Bieniawski (1978)
Serafim and Pereira (1983)
Ironton-Russell
Regression

values. A low-strain modulus derived from downhole seismic measurements was used as a reasonable upper-bound
check on the rock mass modulus. The modulus of intact rock
from laboratory uniaxial compression tests on core samples
is consistent with the observation of Heuze (1980) that field
rock mass modulus values range from 20% to 60% of intact
rock modulus and serve as an additional upper-bound check.

INTERMEDIATE GEOMATERIALS

FIGURE 16 Ratio of rock mass modulus to modulus of intact


rock versus Geological Strength Index (Yang 2006).

A case history described by LaFronz et al. (2003) illustrates


the use of multiple methods for establishing design values of
rock mass modulus. Site characterization for the Colorado
River Bridge (Hoover Dam Bypass Project) included borehole
jack, downhole seismic (compression wave velocity), and laboratory uniaxial compression tests. The major rock unit for the
abutment foundations on the Arizona side of the bridge is
Hoover Dam tuff (welded volcanic ash). Results of field and
laboratory tests used to establish rock mass modulus in the tuff
are summarized in Table 14. Values given for the correlation
with GSI reflect two values of GSI for the tuff, one corresponding to fracture conditions of width = 1 to 5 mm with soft
filling (GSI = 45) and the other corresponding to fracture width
of 0.1 to 1 mm and no filling (GSI = 52). Modulus values based
on downhole p-wave velocities were calculated using equations given by Viskne (1976), described by LaFronz et al.
(2003) as valid at the rock mass scale.
Results were applied as follows. Borehole jack measured
values at stress ranges representative of expected footing
bearing pressures were taken as reasonable values for developing foundation load-deflection curves. Deformation
modulus predicted by the correlation to GSI (Table 12, Hoek
and Brown 1997) provided a cross-check on the borehole
jack measured values. The mean value of modulus from
the borehole jack tests is in the range of the GSI-predicted
TABLE 14
MODULUS VALUES, HOOVER DAM TUFF
(LaFronz et al. 2005)
Method
Borehole jack
Correlation to GSI
Downhole seismic
Uniaxial compression

Mean Modulus (GPa)


2.83
2.34, 3.52
3.31
13.79

A persistent challenge to the geotechnical engineer, and one


that pertains directly to design and construction of drilled
shafts, is defining the boundary between soil and rock. Different approaches to site characterization and evaluation of
geomaterial properties and different design methods are used
when the geomaterial involved is clearly defined as soil or as
rock. However, many geomaterials encountered in practice
exhibit properties that make it difficult to define them clearly
as being soil or rock within the context of standardized classification systems. Geologic processes provide us with a continuum of geomaterial properties and characteristics, some of
which defy simplified categorization.
The term intermediate geomaterial (IGM) has been applied recently to earth materials with properties that are at
the boundary between soil and rock (ONeill et al. 1996).
The criteria are based on (1) whether the material is cohesionless or cohesive and (2) some index of material strength.
Cohesionless IGMs are defined by ONeill et al. (1996) as
very dense granular geomaterials, such as residual, completely
decomposed rock and glacial till, with SPT N60-values
between 50 and 100. Cohesive IGMs are defined as materials
that exhibit unconfined compressive strengths in the range
of 0.5 MPa qu 5 MPa. Specific materials identified
by ONeill et al. (1996) as being cohesive IGMs include
(1) argillaceous geomaterials, such as heavily overconsolidated clays, clay shales, saprolites, and mudstones that are
prone to smearing when drilled; and (2) calcareous rocks such
as limestone and limerock and argillaceous geomaterials that
are not prone to smearing when drilled. The term IGM as used
by ONeill et al. (1996) and subsequently adopted in ONeill
and Reese (1999) and in the draft 2006 Interim AASHTO
LRFD Bridge Design Specifications has been limited specifically to design of drilled shafts and has not been adopted in
the general geotechnical literature. For example, the term
IGM is not used in the FHWA Manual on Subsurface Investigations (Mayne et al. 2001) or in Evaluation of Soil and
Rock Properties, Geotechnical Engineering Circular No. 5
(Sabatini et al. 2002). Responses to Question 8 of the survey
show that most responding states (23) define IGMs for drilled
shaft design according to the criteria of ONeill et al. (1996).
However, six states responded that geomaterials are classified
as either soil or rock and IGM is not used.
According to ONeill and Reese (1999) cohesionless
IGMs may be treated, for practical purposes, in the same

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

30

manner as coarse-grained (cohesionless) soils. They are assumed to respond to loading by rapid dissipation of excess
pore water pressure (fully drained response) and are analyzed
within the context of effective stress. For strength analysis,
cohesionless IGMs are characterized in terms of the effective
stress angle of friction '. It should be noted that some
empirical correlations that apply to cohesionless soils, such
as friction angle estimated from SPT N-values, may not be
applicable to cohesionless IGMs. Specific approaches for
estimating design parameters of shafts in cohesionless IGM
are covered in chapter three.
The definition of cohesive IGMs given earlier is based
on a single index, the unconfined compressive strength. Although this categorization may be useful to identify materials falling into a defined range of intact strength, it does
not necessarily provide the distinction between soil and
rock most relevant to behavior of drilled shafts. To illustrate, consider Figure 17 from Kulhawy and Phoon (1993).
This figure shows the relationship between unit side resistance determined from field load tests on drilled shafts and
one-half of the unconfined compressive strength. Both parameters are normalized by atmospheric pressure pa. Two
categories of load tests were defined; those conducted on
shafts in fine-grained soils (clay) and those in rock. Kulhawy and Phoon relied on the judgment of the original
authors and the database compilers to establish whether the
material was soil or rock. For convenience, the range of
normalized strength that defines cohesive IGM is superimposed on Figure 17. It can be seen that the soil and rock data
constitute apparently different populations, including over
the range of strength that defines cohesive IGM. For purposes of drilled shaft side resistance, therefore, the classification of IGM does not provide a smooth transition from
soil to rock. It may be more meaningful to define the material as being one or the other on the basis of additional
geologic information.

There is no simple answer to the problem of classifying


cohesive materials at the soilrock boundary. Various classifications that distinguish geomaterials on the basis of compressive strength of unweathered rock material are summarized in Figure 18, which includes a proposed classification
by Kulhawy et al. (1991) in which rock strength is defined
relative to that of concrete used in construction, which is assumed to range from 20 kN/m2 (3 ksi) to 100 kN/m2 (15 ksi).
Rock at the high end of the strength scale (>100 kN/m2) is
classified as strong and in most cases would be expected to
be an excellent founding material, except that it would be expensive to excavate. Rock with compressive strength falling
within the range of concrete strength is classified as medium
and the rock mass could be either weaker or stronger than
concrete, depending on weathering and structural features.
For rock classified as weak (<20 kN/m2) foundation capacity
is expected to be governed by the strength of the rock mass.
Materials defined as cohesive IGMs by ONeill et al. (1996)
fall into this strength range. To account properly for the behavior of weak rock in engineered construction, the following additional factors must be considered carefully: geologic
origin, in situ weathering profile, state of stress, groundwater, and construction practices.
A defining characteristic of geomaterials at the soilrock
boundary may be whether or not the in situ material was at one
time rock (geologic origin). This is probably the distinguishing feature between clay and rock in Figure 17. The next geologic consideration is the in situ weathering profile. Igneous,
sedimentary, or metamorphic rocks subjected to in-place
weathering result in geologic profiles that may exhibit the full
range of characteristics, for example, as described in Figure 11
(Key), Sheet 2, under Rock WeatheringAlteration. The
descriptive terms are based on recommendations for describing degree of weathering and alteration by the ISRM. One of
the criteria for distinguishing between residual soil and completely weathered or altered rock is whether the original rock

IGM

FIGURE 17 Side resistance versus geomaterial strength


(Kulhawy and Phoon 1993).

FIGURE 18 Classification for unweathered rock material


strength (Kulhawy et al. 1991).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

31

fabric is still apparent. The highest degree of weathering applies to materials derived from rock but for which the rock
fabric is not apparent. In this case, the material behavior is
controlled by soil fabric and the material should be classified
as residual soil, even though it may contain fragments of
weathered rock. Materials in which the original minerals have
been completely decomposed to secondary minerals but
where the original fabric is intact may exhibit rock material
behavior governed by rock mass features, including both rock
material and discontinuities. The material should be considered to be rock mass, even though it may be highly weathered
or altered and exhibit low compressive strength. Judgment
is always required in assessing whether material behavior is
governed by soil fabric or by rock mass fabric; however, this
is a key factor to be assessed in a design approach. Whether a
geomaterial is assigned the term IGM or weak rock is not
as important as understanding the geologic processes that give
the material its characteristics and engineering properties.

SUMMARY

In this chapter, site characterization methods used to define


subsurface conditions at bridge sites underlain by rock were
reviewed. The survey shows that eight states currently use geophysical methods to determine depth to bedrock and that
seismic refraction is the method used. The literature review
suggests that resistivity methods based on the use of multiple arrays can provide detailed profiles that may be useful
for both design and construction. Karstic areas in limestone
or dolomite terranes with irregular, pinnacled rock surfaces
or solution cavities are examples of sites where recent developments in geophysical methods could be applied.
Every agency responding to the survey uses rock core
drilling as the primary method of subsurface investigation for
rock sockets. Current practice for description and classification of rock core is reviewed. The survey shows that most
states routinely determine the RQD of rock core and that
the uniaxial compressive strength of intact rock (qu) is also
measured by one of the standardized methods. Also from the
survey, it was determined that five states currently classify
all rock mass according to the Geomechanics Classification
System, in which rock mass is assigned a RMR. Twelve
states use RMR occasionally, whereas 14 states indicated
that RMR is never used. Some of the analytical methods developed in recent years and described in subsequent chapters
of this report require the rock mass classification in terms of
RMR. Specifically, RMR can be used to evaluate strength
parameters according to the HoekBrown failure criterion, a
useful approach to quantifying strength of intact or highly
fractured rock masses. RMR is used to establish GSI, which
is required to use the most up-to-date version of the
HoekBrown criterion. RMR and/or GSI are useful for estimating rock mass modulus using the empirical correlations
given in Table 12. The RMR is also recommended in current
FHWA manuals on site characterization and evaluation of

soil and rock properties. Wider use of RMR or GSI classification of rock mass is one way that state DOT agencies can
use the most up-to-date methods for characterizing RMS and
deformation properties.
In situ testing methods that provide information on rock
mass modulus include PMT and borehole jack. Five states
reported using these tests to obtain modulus values for rocksocket design. To use the best available analytical models for
axial and lateral loading, as well as for effective interpretation
of load test results, rock mass modulus is a required parameter.
Currently, it is noted that there is no definitive in situ method
or empirical equation for rock mass modulus that has been calibrated specifically for application to design of rock sockets.
A case history example is presented in this chapter illustrating
the beneficial use of both in situ testing (borehole jack) and
empirical correlations with GSI to establish representative values of rock mass modulus for foundation design.
Site and geomaterial characterization are interrelated with
design, construction, and load testing of drilled shafts in
rock. For design, Figure 3 shows that rock mass engineering
properties required for analysis of rock-socket capacity and
load-deformation response are obtained through field and
laboratory testing. Table 15 is a summary of rock mass characteristics used in design methods for axial and lateral loading. A large X indicates the property is used directly in
design equations that are currently applied widely in practice,
whereas a small x indicates that the characteristic is used
indirectly in the design or that it is required for a proposed
design method that is not used widely. For example, intact rock
modulus ER is not used directly to analyze load-displacement
response of socketed shafts, but may be used to estimate the
rock mass modulus EM, which is used directly in the analytical
equations.
Information obtained through the site investigation
process will be used not only by design engineers but by
contractors who will bid on the work and construct the foundations. As indicated in the flowchart shown in Figure 3,
a goal of site characterization is to obtain information on
constructability. ONeill and Reese (1999) point out that contractors will be most interested in knowing the difficulties
that might be encountered in drilling the rock. Specific information that is useful in assessing the difficulty of drilling
in rock includes loss or gain of drill water; rock type with
lithological description; rock strength; characteristics of
weathering; and rock mass characteristics such as the presence, attitude, and thickness of bedding planes, foliation,
joints, faults, stress cracks, cavities, shear planes, or other
discontinuities. Boring logs, containing most of the information determined by the site investigation, are incorporated
directly into the construction plans by most state DOTs. Any
of the above information not given in the boring logs should
be made available to bidders to facilitate informed decisions.
The same information will be used by the design engineer to
forecast potential construction methods and construction

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

32
TABLE 15
ROCK MASS ENGINEERING PROPERTIES REQUIRED FOR ROCK-SOCKET DESIGN
Design Applications
Lateral Loading

Axial Loading

Rock Mass Characteristic


Compressive strength, intact rock, qu
Split tensile strength, intact rock, qt

Unit Side
Resistance
X

Unit Base
Resistance
X

Axial LoadDisplacement

Ultimate
Resistance
X

Load-Displacement
Continuum
p-y Curve
Methods
Parameter
x

Rock mass strength by Mohr


Coulomb or Hoek and Brown

Shear strength of joint surfaces

Elastic modulus, intact rock, ER

Elastic modulus, rock mass, EM

Rock quality designation (RQD)

Rock Mass Rating (RMR)

Geological Strength Index (GSI)

Notes: X = property is used directly in equations that are currently applied widely in practice.
x = characteristic is used indirectly in the design or it is required for a proposed design method not
widely used.

problems to develop specifications for the project and to


make cost estimates. Rock cores should be photographed
and, when practical, retained for examination by prospective
bidders.
Field load testing, shown in the flowchart of Figure 3
and described in chapter five, provides direct verification of
design assumptions regarding axial and lateral capacity and

load-deformation response. Results of field load testing


also provide the basis for many of the design methods discussed in the next two chapters. For correct interpretation
of load test results, it is imperative that subsurface conditions
and soilrock engineering properties be evaluated as carefully as possible. The properties required for design and
listed in Table 15 are also required for load test interpretation
and for proper extrapolation of load test results to design.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

33

CHAPTER THREE

DESIGN FOR AXIAL LOADING

SCOPE

A rock-socketed drilled shaft foundation must be designed so


that the factored axial resistance is not less than the effects of
the factored axial loads. At the strength limit state, side and
base resistances of the socketed shaft are taken into account.
Design for the service limit state accounts for tolerable
movements of the structure and requires analysis of the axial
load-deformation response of the shaft. In this chapter, current understanding of rock socket response to axial loading
is summarized, based on a literature review. Analysis
methods for predicting axial load capacity and axial loaddisplacement response of shafts in rock and IGM are then
reviewed and evaluated for their applicability to highway
bridge practice.
RELATIONSHIP TO GEOMATERIAL
CHARACTERIZATION

Design for axial loading requires reliable site and geomaterial characterization. Accurate geometric information,
especially depth to rock and thickness of weathered and
unweathered rock layers, is essential for correct analysis of
axial resistance. This information is determined using the
tools and methods outlined in the previous chapter, principally core drilling supplemented by geophysical methods.
Rock mass characterization using the Geomechanics System
(Bieniawski 1989) provides a general framework for assessing the overall quality of the rock mass and its suitability as a
foundation material. Engineering properties of the intact rock
and the rock mass are used directly in the analysis methods
described in this chapter. For example, empirical relationships have been derived between rock-socket unit-side resistance and uniaxial strength of intact rock. Base capacity,
analyzed as a bearing capacity problem, may require uniaxial compressive strength of intact rock, shear strength of
discontinuities, or the HoekBrown strength parameters of
fractured rock mass, depending upon the occurrence, orientation, and condition of joint surfaces in the rock mass below
the base. For analysis of axial load-displacement response,
the rock mass modulus is required. Modulus may be determined from in situ testing, such as pressuremeter or borehole
jack tests, or estimated from rock mass classification parameters as summarized in Table 12. Engineering properties of
rock mass used in conjunction with LRFD methods should
be mean values, not minimum values sometimes used in geotechnical practice.

Several methods proposed in recent years for analysis of


both axial and lateral load response of rock sockets require, as
an input parameter, the GSI proposed by Hoek et al. (1995,
2002). GSI is also correlated to the parameters that establish
the HoekBrown strength criterion for fractured rock masses.
Although GSI is not widely used in foundation engineering
practice at the present time, it likely will become a standard
rock mass characteristic for rock-socket design.
LOAD TRANSFER BEHAVIOR
OF ROCK SOCKETS
Compression Loading

A compressive force applied to the top (head) of a rocksocketed drilled shaft is transferred to the ground through
(1) shearing stress that develops at the concreterock interface along the sides of the shaft and (2) the compressive normal stress that develops at the horizontal interface between
the base of the shaft and the underlying rock. A conceptual
model of the load transfer can be illustrated by considering a
generalized axial load versus displacement curve as shown in
Figure 19 (Carter and Kulhawy 1988). Upon initial loading,
shearing stress develops along the vertical shaftrock interface. For a relatively small load, displacement is small and
the stressstrain behavior at the shaftrock interfaces is
linear (line OA). There is no relative displacement (slip)
between the concrete shaft and surrounding rock and the system may be modeled as being linearly elastic. With increasing load, the shear strength along some portion of the shaft
sidewall is exceeded, initiating rupture of the bond and relative slip at the shaftrock interface. The load-displacement
curve becomes nonlinear as rupture, and slip progress and a
greater proportion of the applied load is transferred to the
base (line AB). At some point, the full side resistance is
mobilized, and there is slip along the entire surface (full
slip condition), and a greater proportion of the applied load
is transferred to the shaft base (beyond point B in Figure 19).
If loading is continued to a displacement sufficient to cause
failure of the rock mass beneath the base, a peak compressive
load may be reached. In practice, design of drilled shafts
in rock requires consideration of (1) deformation limits and
(2) geotechnical and structural capacity (strength limit
states). Geotechnical capacity in compression is evaluated in
terms of limiting side and base resistances. Load transfer
in uplift involves the same mechanisms of side resistance
mobilization as described previously for compression.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

34

asperities from the direction of shear displacement. The interface shear strength () is then given by

Load, Qc

= c + n tan ( + )
Full slip

B
Progressive slip

A
Linear elastic

Settlement, wc
FIGURE 19 Idealized load-displacement behavior.

A rigorous model for the behavior of a rock-socketed


drilled shaft under axial compression would provide a prediction of the complete load-displacement curve. In reality,
the mechanisms of side and base load transfer are complex
and can only be modeled accurately through the use of sophisticated numerical methods, such as finite-element or
boundary-element methods. Input parameters required for
accurate modeling are not normally available for design. In
recent years, several researchers have presented simplified
methods of analysis that provide bounds on the expected and
observed behaviors for shafts that fall within the range of
conditions typically encountered in practice. Methods most
relevant to rock-socketed bridge foundations are presented in
this chapter. Some of the more important behavioral aspects
pertaining to side and base resistance and their mobilization
are described first.
Side Resistance Mechanisms

The conditions of the sidewall interface determine the


strength and load transfer in side resistance. Side resistance
often exhibits a bond component that may exist physically
as a result of the cementation between the concrete and rock
and from mechanical interlocking between asperities along
the interface. If the shearing strength of the interface is modeled as a MohrCoulomb material, the bond component can
be considered as the interface adhesion, c. If displacements
are sufficient, the interface bond is ruptured and the cohesion
component of resistance may be diminished. The second
mechanism of resistance is frictional. Physically, the frictional resistance can have two components. The first is the
sliding friction angle of the interface, . The second is
mechanical dilatancy, which can be described as an increase
in the interface normal stress in response to the normal
displacement (dilation) required to accommodate shear displacement of a rough surface. For mathematical simplicity,
dilatancy can be quantified in terms of the angle of dilation
(), where corresponds to the average angle of triangular

(18)

in which n = interface normal stress. Physically, all three


components of strength (c, , ) may vary with displacement. The initial shear strength may have both cohesive and
frictional components. Following rupture, the cohesion is
probably decreased and dilation is mobilized. With further
displacement, dilation may cease and resistance may be
purely frictional and correspond to the residual friction angle.
In addition, field conditions of construction can significantly
affect the nature of the sidewall interface and, in practice,
will determine the relative contributions of cohesion, friction, and dilatancy to shearing resistance. For example, the
bond (adhesion) may be partially or completely prevented
by the presence of drilling slurry, or by smearing, which
occurs in some argillaceous rocks or in rocks that are sensitive to property changes in the presence of water. Dilatancy
is a function of interface roughness and shear strength of
the intact rock forming the asperities. Sidewall roughness is
determined in part by rock type and texture, but can also be
affected by construction tools and practices. Practices that
result in a smooth sidewall will reduce dilatancy compared
with practices that provide a rough sidewall (Williams and
Pells 1981; Horvath et al. 1983).
Johnston and Lam (1989) made detailed investigations of
the rockconcrete interface with the goal of better understanding the factors that determine interface roughness and
its influence on side-load transfer. Figure 20a shows an idealized section of a rock socket following construction. An
initial normal force exists between the rock and concrete.
When the shaft is loaded vertically, the shearing resistance
develops and the rock mass will deform elastically until slip
occurs. Figure 20a and b show the positions of the shaft
before and after slip displacement. These two conditions are
represented by 2-D models in Figure 20c and d, respectively.
Figure 20d illustrates the dilation that occurs as a result of geometrical constraints. Dilation occurs against the surrounding
rock mass, which must deform to compensate for the increase
in socket diameter, resulting in an increase in the interface
normal stress. The average normal stress increase (n) can be
approximated using the theoretical solution that describes
expansion of an infinite cylindrical cavity, as follows:
n =

EM r
1+ r

(19)

where EM and are the rock mass modulus and Poissons ratio, respectively; r is the dilation, and r is the original shaft
radius. A normal stiffness K can be defined as the ratio of
normal stress increase to dilation, as follows:
K=

EM
n
=
r r (1 + )

Copyright National Academy of Sciences. All rights reserved.

(20)

Rock-Socketed Shafts for Highway Structure Foundations

35

FIGURE 21 Theoretical base load transfer (Rowe and


Armitage 1987b).

of modulus ratio is more significant at lower embedment


ratios and, in general, base load transfer increases with increasing modulus ratio. Cases that result in the most base
load transfer correspond to low embedment ratio with high
modulus ratio (shaft is rigid compared to rock mass);
whereas the smallest base load transfer occurs at higher embedment ratios and low modulus ratio (stiff rock mass).

Shaft Geometry and Relative Rigidity

The proportion of load transferred to the base will also


vary with the stiffness of the rock mass beneath the base of
the shaft relative to the stiffness of the rock along the side. In
many situations, a rock socket is constructed so that the base
elevation corresponds to relatively sound or intact rock,
and it may be necessary to excavate through weathered or
fractured rock to reach the base elevation. In that case, the
modulus of the rock mass below the base may be greater than
that of the rock along the sidewall of the socket. Osterberg
and Gill (1973) demonstrate the difference in load transfer in
side and base resistances for two conditions, one in which the
base modulus is twice that of the sidewall rock modulus and
one where the base rock has a much lower modulus than that
of the rock surrounding the shaft side. Their results show
that base load transfer increases as the ratio Eb/Er increases
(Figure 22).

Load transfer in a rock socket depends on the geometry,


expressed by the embedment ratio (depth/diameter), and the
stiffness of the concrete shaft relative to stiffness of the surrounding rock mass. Figure 21, based on finite-element
analysis, illustrates this behavior for the initial (no slip) part
of the load-displacement curve. In Figure 21, L = socket
length, D = shaft diameter, Ep = modulus of the shaft, Er =
modulus of rock mass above the base, Eb = modulus of rock
mass below the base, Qb = load transmitted to the base, and
Qt = load applied to the head of the shaft. The portion of applied axial compressive load that is transferred to the base is
shown as a function of embedment ratio and modulus ratio.
With increasing embedment ratio, the relative base load
transfer decreases. For embedment ratios of 10, less than
10% of the applied load is transferred to the base. The effect

Load transfer is affected significantly by the roughness of


the sidewall interface. Fundamentally, this can be explained
by the higher load transfer in side shear reducing the proportion of load transferred to the base. Because side resistance
increases with interface roughness, rock sockets with higher
interface roughness will transfer a higher proportion of load
in side resistance than smooth sockets. The complex interrelationships between load transfer, interface roughness, modulus ratio, and embedment ratio have been studied by several
researchers, and the reader is referred to Pells et al. (1980),
Williams et al. (1980), Rowe and Armitage (1987a), and
Seidel and Collingwood (2001) for more detailed discussions. Six state DOTs indicated the use of grooving tools or
other methods to artificially roughen the sidewalls of rocksocketed shafts.

FIGURE 20 Idealized rockconcrete interface under axial


loading (Johnston and Lam 1989).

Assuming the deformation r is small compared with r, and


EM and v can be considered to be constant for the stress range
considered, it follows that the behavior of the rockconcrete
interface is governed by CNS conditions. This concept forms
the basis of the CNS direct shear test, in which the normal
force is applied through a spring (Johnston et al. 1987; Ooi
and Carter 1987).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

36

FIGURE 22 Effect of rock mass modulus at base on axial load


transfer (Wyllie 1999, based on Osterberg and Gill 1973).
Base Condition

In many cases encountered in practice there is uncertainty


about conditions at the base of the shaft. Most transportation agencies include, in their drilled shaft specifications,
limits on the amount of drill cuttings, water, or slurry that is
permissible at the base before concrete placement (Survey
Question 34). However, compliance is not always verified
and in some cases there is a perception that it is not practical
to clean or inspect the base of the socket. In these cases, the
designer may assume that base resistance will not develop
without large downward displacement and for this reason
base resistance is sometimes neglected for design purposes.
Ten states indicated in their responses to Question 14 of
the survey that rock-socketed shafts are sometimes designed
under the assumption of side resistance only. The draft
Interim 2006 AASHTO LRFD Bridge Design Specifications
state that Design based on side-wall shear alone should
be considered for cases in which the drilled hole cannot
be cleaned and inspected or where it is determined that
large movements of the shaft would be required to mobilize
resistance in end bearing. Table 16 lists the most common
reasons cited by foundation designers for neglecting base
resistance in design, along with actions that can be taken to
address the concern.
Crapps and Schmertmann (2002) suggest that accounting
for base resistance in design and using appropriate construc-

tion and inspection techniques to ensure quality base conditions is a better approach than neglecting base resistance. The
authors support their recommendations with field load test
results in which load transferred to the base was measured.
The database consisted of 50 Osterberg load cell (O-cell)
tests and 22 compression tests in which the load was applied
to the top of the shaft. Of those, 30 of the O-cell tests and 4
of the top load tests were conducted on rock-socketed shafts.
Eight of the O-cell tests (27%) showed evidence of bottom
disturbance in the O-cell load-displacement curves. Results
from the 34 tests are plotted in Figure 23 in terms of base load
ratio (Qb = base load, Qt = actual top load or top load inferred
from the O-cell test) versus socket-effective depth-to-diameter
ratio (L/B). For some of the shafts, multiple measurements
are included at different values of load and displacement.
However, all of the base load ratio values correspond to
downward displacements at the top of the shaft that range
from 2.5 mm to 25.4 mm, with most in the range of from 3
to 15 mm. These values are within the service limit state for
most bridge foundations. Additional details regarding the test
shafts, subsurface profiles, and load test interpretation are
given in Crapps and Schmertmann (2002).
Several important observations arise from the data shown in
Figure 23. First, base resistance mobilization represents a significant contribution to overall shaft resistance at downward
displacements corresponding to typical service loads. Second,
the magnitude of base resistance is generally greater than predicted by elasticity-based numerical solutions (e.g., compare
with Figure 21). The dashed lines in Figure 23 represent approximate upper and lower bounds to the data from top load
tests and O-cell tests without bottom disturbance. For the most
part, O-cell tests that exhibited bottom disturbance fall below
the lower-bound curve. Although the data are not sufficient to
provide design values of base load transfer in advance for a
given situation, they provide compelling evidence that shaft
design in rock should account properly for base resistance, and
that quality construction and inspection aimed at minimizing
base disturbance can provide performance benefits.

Time Dependency

Time-dependent changes in load transfer may occur in rocksocketed shafts under service load conditions. Ladanyi (1977)
reported a case in which the bearing stress at the base of an instrumented rock socket increased, at a steadily decreasing
rate, over a period of 4 years; although the total applied head

TABLE 16
REASONS FOR NEGLECTING BASE RESISTANCE AND CORRECTIVE ACTIONS (after
Crapps and Schmertmann 2002)
Reason Cited for Neglecting Base Resistance
Settled slurry suspension
Reluctance to inspect bottom
Concern for underlying cavities
Unknown or uncertain base resistance

Correction
Utilize available construction and inspection methods
Utilize available construction and inspection methods
Additional inspection below base
Load testing

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

37
1
0.9
0.8
0.7

Qb/Qt

Qs =

surface

O-cell, good base


O-cell, base disturbed
top load test

0.5
0.4
0.3
0.2
0.1
0
1

(21)

in which Qs = total side resistance (force), fsu = unit side resistance (stress), A = surface area along the side of the socket, B =
socket diameter, and L = socket length. In practice, socket side
resistance capacity is calculated by assuming that a single average value of unit side resistance acts along the concreterock
interface, for each rock layer. This value of f is multiplied by
the area of the interface to obtain total side resistance Qs, or

0.6

fsu dA = B fdz

10

Qs = fsu BL

(22)

L/B

FIGURE 23 Base load transfer interpreted from load tests


(data from Crapps and Schmertmann 2002).

load remained essentially constant. However, the percentage


of total load carried by the base after this period was still less
than 10% of the applied load and agreed quite well with predictions based on elastic theory. Tang et al. (1994) described
a similar monitoring program on a shaft socketed into karstic
dolomite supporting a building on the University of Tennessee campus. Some change in load transfer occurred following the end of construction; however, most of the change
was from side resistance in the overlying soil (decreased) to
side resistance in the rock socket (increased). Neither case
would suggest changes in design of rock sockets to account
for the time dependency of load transfer mechanisms.

CAPACITY UNDER AXIAL LOADING

The factored axial resistance of a drilled shaft in compression is the sum of the factored side resistance and the factored base resistance. The factored resistances are calculated by multiplying appropriate resistance factors by the
nominal resistances, which are generally taken as the ultimate values. One approach to the design process depicted
in Figure 3 of chapter one is to size the foundation initially
to achieve a factored resistance that exceeds the factored
loads. The trial design is then analyzed to predict loaddisplacement response. If necessary, revised trial dimensions can then be analyzed until all of the design criteria are
satisfied, including the movement criteria associated with
the service limit state. In the case of axial loading, the ultimate
side and base resistances are required to establish the initial
trial design.

Side Resistance

The ultimate side resistance of a rock socket is the summation of peak shearing stress acting over the surface of the
socket, expressed mathematically by

Methods for predicting socket side resistance are, therefore,


focused on the parameter fsu.
The interaction between a rock mass and drilled shaft that
determines side resistance is complex. The principal factors
controlling this interaction include:
Rock material strength;
Rock mass structure (discontinuities);
Modulus of the concrete relative to modulus of the rock
mass;
Shear strength mobilized by dilatancy;
Confining stress; and
Construction-related factors, including roughness of
shaftrock interface.
Geomechanical models that account for these factors (to
varying degrees) are described in the literature (e.g., Rowe
and Pells 1980); however, the methods required to obtain the
necessary input parameters normally fall outside the scope of
a typical investigation conducted for the design of highway
bridge foundations. More realistically, methods based on the
strength of intact rock, in some cases with modifications to
account for one or more of the other factors, have been used
successfully and are more rational than some of the strictly
empirical methods or presumptive values. This approach represents a practical compromise between oversimplified empirical methods and more sophisticated numerical methods
that might be warranted only on larger projects. The methods
are summarized here.
Methods Based on Rock Compressive Strength
A practical approach to evaluating average unit side resistance is to relate fsu to the strength of the intact rock material.
The rock material strength parameter most often measured is
the uniaxial compressive strength (qu). In this approach, values of fsu are determined from full-scale field load tests in
which ultimate side resistance (Qs) has been determined.
This value is divided by the socket side area (As) to obtain an
average value of unit side resistance at failure:
fsu = Qs/As

Copyright National Academy of Sciences. All rights reserved.

(23)

Rock-Socketed Shafts for Highway Structure Foundations

38

Early studies relating fsu to qu include those by Rosenberg


and Journeaux (1976) and Horvath (1978, 1982). Other researchers have continued to expand the available database
and propose equations relating unit side resistance to rock
strength on the basis of statistical best-fit analyses. Notable
studies include those of Williams and Pells (1981), Rowe and
Armitage (1984, 1987b), Bloomquist and Townsend (1991),
McVay et al. (1992), and Kulhawy and Phoon (1993). These
studies, including the proposed equations relating unit side
resistance to rock strength are reviewed briefly.
Horvath and Kenney (1979) proposed the following correlation between side resistance and compressive strength:
fsu = b qu

(24)

in which fsu = ultimate unit side resistance, qu = compressive


strength of the weaker material (rock or concrete), and where
b ranges from 0.2 (smooth) to 0.3 (rough). Both fsu and qu in
Eq. 24 are in units of MPa. Eq. 24 can be expressed in normalized form by dividing both unit side resistance and compressive strength by atmospheric pressure (pa = 0.1013 MPa).
This results in the following expression, which is equivalent
to Eq. 24 with b = 0.2:
fsu
q
= 0.65 u
pa
pa

(25)

A modified relationship was recommended by Horvath et al.


(1983) to account for shafts with artificially roughened
(grooved) sockets. The suggested relationship is given by
fsu = 0.8 [ RF ]0.45 qu
RF =

(26)

rh Lt
rs Ls

(27)

in which RF = roughness factor, rh = average height of asperities, rs = nominal socket radius, Ls = nominal socket
length, and Lt = total travel distance along the socket wall
profile. A device (caliper) was used to measure field roughness for determination of the parameters needed in Eq. 27, as
described by Horvath et al. (1993).
The FHWA Drilled Shaft Manual (ONeill and Reese 1999)
and the draft 2006 Interim AASHTO LRFD Bridge Design
Specifications have adopted Eqs. 24 to 27 as a recommended
method for selection of design side resistance for shafts in rock.

AASHTO refers to this as the Horvath & Kenney method. A


socket that is not specified to be artificially roughened by
grooving is considered smooth and side resistance is governed by Eq. 25. If the socket is artificially roughened, Eq. 26
is recommended; however, this requires estimation or measurement of roughness as defined by Eq. 27.
Rowe and Armitage (1987b) summarized the available
data on side resistance of rock sockets, including the databases used by Williams et al. (1980), Williams and Pells
(1981), and Horvath (1982). The suggested correlation for
regular clean sockets, defined as roughness classes R1, R2,
and R3 in Table 17, is given as
fsu = 0.45 qu

To account for rough sockets, defined as category R4, side


resistance is increased and the following is recommended:
fsu = 0.6 qu

(29)

The correlation suggested by Horvath and Kenney (1979)


as given by Eq. 25 represents a lower bound to the data used
by Rowe and Armitage (1987b)
Kulhawy and Phoon (1993) incorporated the database
compiled by Rowe and Armitage, which included more than
80 load tests from more than 20 sites, and the data reported
by Bloomquist and Townsend (1991) and McVay et al.
(1992) consisting of 47 load tests to failure from 23 different
Florida limestone sites. Linear regression was conducted on
two sets of data, one consisting of all data points and the
other using data that were averaged on a per-site basis. Averaging eliminates the bias associated with multiple load
tests conducted at many of the sites. All stresses are normalized by atmospheric pressure, pa (0.1013 MN/m2) and normalized values of one-half of the uniaxial compressive
strength were plotted against normalized values of average
unit side resistance on a loglog plot as shown in Figure 24
(shown previously as Figure 17).
The following exponential expression provides a best-fit
to the available data for rock:
fsu
qu
=C
2 pa
pa

TABLE 17
SHAFT ROUGHNESS CLASSIFICATION (after Pells et al. 1980)
Roughness
Class
R1
R2
R3
R4

(28)

Description
Straight, smooth-sided socket; grooves or indentations less than 1 mm deep
Grooves 14 mm deep, >2 mm wide, spacing 50200 mm
Grooves 410 mm deep, >5 mm wide, spacing 50200 mm
Grooves or undulations >10 mm deep, >10 mm wide, spacing 50200 mm

Copyright National Academy of Sciences. All rights reserved.

(30)

Rock-Socketed Shafts for Highway Structure Foundations

39

capture all of the mechanisms affecting unit side resistance


and because the database incorporates load test results for
many different rock types. Use of these empirical correlations for design should therefore be conservative (i.e., C = 1)
unless a site-specific correlation has been developed that justifies higher values. Research on development of methods
that account for these additional factors affecting peak side
resistance is summarized here.

Methods Based on Additional Rock


Mass Parameters

FIGURE 24 Unit side resistance versus strength (Kulhawy and


Phoon 1993).

where values of the coefficient C = 1 represents a reasonable


lower bound, C = 2 represents the mean behavior, and C = 3
corresponds to an upper bound for artificially roughened
sockets. Kulhawy and Phoon (1993) noted that the expressions and variations are consistent with those reported by
Rowe and Armitage (1987b). Values of fsu obtained using
Eq. 30 with C = 2 (mean) are identical to the values obtained
using Eq. 28 (Rowe and Armitage) corresponding to the mean
trend for smooth sockets. The expression of Rowe and Armitage for rough sockets (Eq. 29) corresponds to Eq. 30 with
C = 2.7.
Kulhawy et al. (2005) recently reexamined the data available and attempted to evaluate them in a more consistent
manner. Only data showing load-displacement curves to failure were incorporated into the analysis, so that the interpreted failure load could be established in a consistent manner. Based on the updated analysis, the authors recommend
the use of Eq. 30 with C = 1 for predicting side resistance of
normal rock sockets for drilled shafts. The authors also note
the importance of using compressive strength values (qu)
obtained from laboratory uniaxial compression tests, not
from point load tests.

Williams et al. (1980) and Rowe and Armitage (1987a) point


out that unit side resistance determined strictly by empirical
correlations with uniaxial compressive strength does not account explicitly for the degree of jointing in the rock mass.
Figure 25 shows the potential influence on average side resistance (denoted by in the figure) of jointing in terms of
the ratio of rock mass modulus to modulus of intact rock.
Both theoretical and experimental curves show substantially
reduced side resistance for rock that may have high intact
strength and stiffness but low mass modulus.
The draft Interim 2006 AASHTO LRFD Bridge Design
Specifications adopt the Horvath and Kenney method described earlier (Eqs. 2527), with the following modification
as recommended ONeill and Reese (1999). Values of unit
side resistance calculated by either Eq. 25 or Eq. 26 are modified to account for rock mass behavior in terms of RQD,
modulus ratio (EM/ER), and joint condition using the factor
as defined in Table 18. In Table 18, fdes is the reduced unit

In summary, Eq. 30 with C = 1 provides a conservative


estimate of design ultimate side resistance, based on the most
up-to-date analysis of the available data. Use of C values
greater than 1 for design should be verified by previous experience or load testing. Load test results that exhibit values
of C in the range of 2.7 to 3 demonstrate the potential increase in side resistance that is possible if the sidewalls of the
socket are roughened. These upper-bound values of fsu should
only be considered when they can be validated by field load
testing.
The methods described rely on empirical relationships
between side resistance and a single parameter, uniaxial compressive strength, to represent the rock mass. There is significant scatter in the database because the relationships do not

FIGURE 25 Effect of rock mass modulus on average unit side


resistance (Rowe and Armitage 1987a).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

40
TABLE 18
SIDE RESISTANCE
REDUCTION BASED ON
MODULUS REDUCTION
(ONeill and Reese 1999)
EM/ER
1.0
0.5
0.3
0.1
0.05

= fdes/fsu
1.0
0.8
0.7
0.55
0.45

side resistance recommended for design. The modulus reduction ratio (EM/ER) is given in Table 13, in chapter two,
based on RQD. However, application of the -factor may be
questionable because the RQD and rock mass modulus were
not accounted for explicitly in the original correlation analysis by Horvath and Kenney (1979). Because the load test data
included sites with RQD less than 100 and modulus ratio values less than one, it would appear that these factors affected
the load test results and are therefore already incorporated
into the resulting correlation equations.
Interface roughness is identified by all researchers as having a significant effect on peak side resistance. Pells et al.
(1980) proposed the roughness classification that assigns a
rock socket to one of the categories R1 through R4 as defined
in Table 17. The criteria are based on observations of sockets drilled in Sydney sandstone and the classification reportedly forms the basis for current practice in that city (Seidel
and Collingwood 2001). The Rowe and Armitage (1987b)
correlation equations were developed by distinguishing between roughness classes R1R3 (Eq. 28) and roughness class
R4 (Eq. 29). Horvath et al. (1983) proposed the roughness
factor (RF) defined in Eqs. 26 and 27 as presented earlier.
Despite these efforts, selection of side resistance for rocksocket design in the United States is done mostly without
considering interface roughness explicitly.
Seidel and Haberfield (1994) developed a theoretical model
of interface roughness that accounts for the behavior and characteristics of socket interfaces under CNS conditions. Roughness is modeled using a quasi-probabilistic approach that
involves fractal geometry to predict the distribution and characteristics of asperities. Results of the interface model and
laboratory CNS testing are incorporated into the computer program ROCKET that predicts the axial load-displacement curve,
including post-peak behavior. Extending this work, Seidel and
Collingwood (2001) proposed a nondimensional parameter
defined as the shaft resistance coefficient (SRC) to account for
the factors that influence side resistance, as follows:
SRC = c

n r
1 + ds

fsu = (SRC) qu

(31)
(32)

in which c = construction method reduction factor, as defined in Table 19; n = ratio of rock mass modulus to uniaxial
compressive strength of intact rock (EM/qu); v = Poissons
ratio; r = mean roughness height; and ds = socket diameter.
Implementation of the SRC in design requires an estimate
of socket roughness in terms of r. As noted by the authors,
reliable measurements of roughness are not undertaken in
routine design. However, the SRC factor incorporates many
of the significant parameters that influence side resistance,
including rock mass modulus, Poissons ratio, and intact rock
strength, and provides a framework for taking into account
socket roughness and construction effects.
The SRC method represents the type of approach that
holds promise for improved methods for selecting design side
resistance. Although more detailed guidance is required for
determination of socket roughness and construction effects,
improvements in reliability of design equations are possible
only if the relevant factors controlling side resistance are incorporated properly. Advancement of the SRC or other robust
methods can be facilitated by promoting the awareness of
engineers involved in field load testing of the importance of
collecting appropriate data on rock mass characteristics. Documentation of RMS and modulus along with careful observation
and documentation of construction procedures would allow
these methods to be evaluated against load test results. The key
parameter that is currently missing from the database is socket
roughness. ONeill et al. (1996) point out that roughness can be
quantified approximately by making electronic or mechanical
caliper logs of the borehole, and that such borehole calipers are
available commercially. Seidel and Collingwood (2001) describe a device called the Socket-Pro that is operated remotely
and records sidewall roughness to depths of 60 m.

Geomaterial-Specific Correlations
Correlations between unit side resistance and intact rock
strength that are based on a global database (e.g., Figure 24,
Eq. 30) exhibit scatter and uncertainty because the results
reflect the variations in interface shear strength of different
rock types, interface roughness, and other factors that control
side resistance. For this reason, selection of design side resistance values based on such correlations should be considered as first-order estimates and the philosophy underlying
their use for design is that a lower-bound, conservative relationship should be used (e.g., C = 1 in Eq. 30). Alternatively,
correlations have been developed for specific geomaterials.
Correlations identified by the literature review and the survey
are summarized here.
Florida Limestone
Limestone formations in Florida are
characterized by highly variable strength profiles, the presence of cavities that may be filled with soil, and interbedding
of limestone with sand and marine clay layers (Crapps 1986).
Locally, geotechnical engineers distinguish between lime-

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

41
TABLE 19
CONSTRUCTION METHOD REDUCTION FACTORS, c (Seidel and Collingwood 2001)
Construction Method
Construction without drilling fluid
Best practice construction and high level of construction control (e.g., socket
sidewalls free of smear and remolded rock)
Poor construction practice or low-quality construction control (e.g., smear or
remolded rock present on rock sidewalls)
Construction under bentonite slurry
Best practice construction and high level of construction control
Poor construction practice or low level of construction control
Construction under polymer slurry
Best practice construction and high level of construction control
Poor construction practice or low level of construction control

rock and limestone; the former defined informally as


material with qu less than approximately 13.8 MN/m2 (2,000
psi). McVay et al. (1992) conducted a study of design methods used to predict unit side resistance of drilled shafts in
Florida limestone. Based on a parametric finite-element
study and a database of 14 case histories consisting of fullscale load tests and field pullout tests, the following expression was found to provide a reasonable estimate of ultimate
unit side resistance:
fsu =

1
qu qt
2

(33)

In which qu = uniaxial compressive strength and qt = split tensile strength. To account for the effect of material strength
variability on side resistance, the authors recommend a minimum of 10 (preferably more) core samples be tested in unconfined compression and splitting tensile tests. The mean
values of qu and qt are used in Eq. 33. The standard error
of the mean from the laboratory strength tests can be used to
estimate the expected variation from the mean side resistance, for a specified confidence level.
According to Lai (1998), design practice by the Florida
DOT is based on a modified version of the McVay et al.
relationship in which spatial variations in rock quality are
incorporated by multiplying the unit side resistance, according to Eq. 33, by the average percent recovery (REC) of rock
core expressed as a decimal, or:

( fsu )design =

REC(%) 1
qu qt
100% 2

c
1.0
0.30.9
0.70.9
0.30.6
0.91.0
0.8

used for design should be limited to the design strength of the


shaft concrete.
Side resistance values in Florida limestone have also been
evaluated using a small-scale field pullout test devised by
Schmertmann (1977) for the Florida DOT and shown
schematically in Figure 26. A grout plug is placed into a 140mm-diameter cored hole at the bottom of a 165-mm-diameter
hole drilled to the test depth in rock. Overburden soils are
supported by a 200-mm-diameter casing. The grout plug is
reinforced with a wire cage and a threaded high-strength steel
bar extends from the bottom of the plug to the ground surface. A center hole jack is used to apply a pullout force to the
bar. The grout plug is typically 610 mm (2 ft) in length, but
other lengths are also used. The average unit side resistance
is taken as the measured pullout force divided by the sidewall
interface area of the plug (Eq. 23). Results of pullout tests
were included in the database of McVay et al. (1992) that
forms the basis of Eq. 33, and McVay et al. recommend the
test as an alternative method for estimating side resistance for
design.
Cohesionless IGM
The FHWA Drilled Shaft Manual
(ONeill and Reese 1999) recommends a procedure for calculating unit side resistance specifically for cohesionless
IGM. These are granular materials exhibiting SPT N60-values
between 50 and 100. The method follows the general
approach for calculating side resistance of drilled shafts in
granular soils, given by
fsu = v' K o tan '

(34)

Lai (1998) also recommends using larger diameter doubletube core barrels (61 mm to 101.6 mm inner diameter) for obtaining samples of sufficient quality for laboratory strength
tests. Analysis of the laboratory strength data involves discarding all data points above or below one standard deviation
about the mean, then using the mean of the remaining values
as input to Eq. 34. Crapps (2001) recommends using RQD in
place of REC in Eq. 34 and points out that values of qu and qt

(35)

in which fsu = ultimate unit side resistance, v' = vertical effective stress, Ko = in situ coefficient of lateral earth pressure, and
' = effective stress friction angle of the IGM. The modifications to account for cohesionless IGM behavior are incorporated into empirical correlations with the N-value as follows:
'p = 0.2 N 60 pa
OCR =

'p
v'

Copyright National Academy of Sciences. All rights reserved.

(36)
(37)

Rock-Socketed Shafts for Highway Structure Foundations

42
nut
plate
centerhole
jack

steel bearing
plate

timber

35 mm dia
threaded bar

200-mm dia
casing

top of rock

drilled
hole,
165 mm dia

FIGURE 27 Factor for cohesive IGM (ONeill et al. 1996).

0.76 m cored
hole, 140 mm dia

0.61 m grout plug


w/ #9 wire cage

FIGURE 26 Small-scale pullout test used in Florida limestone


(after Crapps 1986).

different value of interface friction applies, then the parameter can be adjusted by
= 26

N
60

' = tan 1
12.2 + 20.3 'v
p

a
K o = (1 sin ')OCR sin '

0.34

(38)

(39)

where p' = preconsolidation stress, 'v = average vertical


effective stress over the layer, N60 = SPT N-value corresponding to 60% hammer efficiency, and OCR = overconsolidation ratio. ONeill et al. (1996) reported good agreement
with results of load tests on shafts in residual micaceous sands
in the Piedmont province (Harris and Mayne 1994) and granular glacial till in the northeastern United States.
Soft Argillaceous Rock
Side resistance in weak shales and
claystones can be approximated for design using the relationships given previously for rock. Alternatively, a procedure
for evaluating unit side resistance specifically in argillaceous
(containing clay) cohesive IGMs is presented in the FHWA
Drilled Shaft Manual (ONeill and Reese 1999). The ultimate unit side resistance is given by
fsu = qu

(40)

where qu = compressive strength of intact rock and = empirical factor given in Figure 27. In Figure 27, n = fluid pressure exerted by the concrete at the time of the pour and p =
atmospheric pressure in the same units as n. As indicated
in Figure 27, the method is based on an assumed value of
interface friction angle rc = 30 degrees. If it is known that a

tan rc
tan 30 o

(41)

The method is based on work reported by Hassan et al.


(1997) in which detailed modeling and field testing were
conducted to study side resistance of shafts in clayshales
of Texas. ONeill and Reese (1999) provide additional equations for modifying side resistance for roughness, the presence of soft seams, and other factors, and the reader is
advised to consult their work for these additional details.
Correlations with In Situ Tests
The survey responses indicate that several states use measurements from field penetration tests to estimate unit side
resistance in weak rock. As an example, consider the Texas
cone penetration test (TCPT) described in chapter two. The
Texas DOT Geotechnical Manual, which is accessible online (2005), presents graphs for estimating design values of
allowable unit side and base resistances as a function of the
penetration resistance (millimeters of penetration per 100
blows), for materials that exhibit TCPT blowcounts of
greater than 100 blows per 300 mm. For materials exhibiting
fewer than 100 blowcounts, separate graphs are provided for
allowable values of unit side and base resistances. The Missouri DOT also reports using the TCPT to correlate allowable side and base resistances in weak rock.
Base Resistance

Load transmitted to the base of a rock-socketed shaft, expressed as a percentage of the axial compression load applied

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

43
Rock Mass Conditions
Joint Dip
Angle, from
horizontal

Joint
Spacing

Failure

Illustration

Mode
(a) Brittle Rock:
Local shear failure caused
by localized brittle fracture

INTACT / MASSIVE

N/A

S>>B

(b) Ductile Rock:


General shear failure along
well-defined shear surfaces

(c) Open Joints:


Compression failure of
individual rock columns

STEEPLY DIPPING JOINTS

at the head, can vary over a wide range at typical working


loads. Several authors suggest a typical range of 10% to 20%
of the head load (Williams et al. 1980; Carter and Kulhawy
1988), and some authors suggest that base resistance should
be neglected entirely for rock-socket design (Amir 1986).
Elasticity solutions show that base load transfer depends on
the embedment ratio (L/B) and the modulus ratio (Ec/Er). The
ratio of base load to applied load (Qb/Qc) decreases with increasing L/B (see Figure 21) and increases with increasing
modular ratio. As discussed previously, there is ample evidence that base resistance should not be discounted in most
cases (Figure 23), and that construction and inspection methods are available to control base quality. Load tests, described
in chapter five, provide a means to determine the effects of
construction on base load transfer.

s
S<B

(d) Closed Joints


General shear failure along
well defined failure
surfaces; near vertical
joints

70 < < 90

S>B

20 < < 70

S < B or
S > B if
failure
we dge can
de ve lop
along joints
H

Analytical solutions for bearing capacity of rock are based


on the general bearing capacity equation developed for soil,
with appropriate modifications to account for rock mass
characteristics such as spacing and orientation of discontinuities, condition of the discontinuities, and strength of the rock
mass. Typical failure modes for foundations bearing on rock
are shown in Figure 28. The failure modes depicted were intended to address shallow foundations bearing on rock (Sowers 1976); however, the general concepts should be applicable to bearing capacity of deep foundations. The cases shown
can be placed into four categories: massive, jointed, layered,
and fractured rock.

Massive Rock
For this case, the ultimate bearing capacity will be limited to
the bearing stress that causes fracturing in the rock. An intact
rock mass can be defined, for purposes of bearing capacity
analysis, as one for which the effects of discontinuities are
insignificant. Practically, if joint spacing is more than four
to five times the shaft diameter, the rock is massive. If the
base is embedded in rock to a depth of at least one diameter, the failure mode is expected to be by punching shear
(Figure 28, mode a). In this case, Rowe and Armitage
(1987b) stated that rock fracturing can be expected to occur
when the bearing stress is approximately 2.7 times the rock
uniaxial compressive strength. For design, the following is
recommended:
qult = 2.5 qu

LAYERED

(42)

rigid

Eq. 43

Eq. 44

Eqs. 4552

(e) Open or Closed Joints:


Failure initiated by splitting
leading to general shear
failure; near vertical joints

Eqs. 5354

(f) General shear failure


with potential for failure
along joints; moderately
dipping joint sets.

Eqs. 4552

0 < < 20

Limiting
va lue of H
w/re to B is
dependent
upon
material
properties

weak, compressible

H
rigid

(g) Rigid layer over weak


compressible layer:
Failure is initiated by
tensile failure caused by
flexure of rigid upper layer
(h) Thin rigid layer over
weak compressible layer:
Failure is by punching
shear through upper layer

N/A

N/A

weak, compressible

FRACTURED

Qb = qult Ab = qult [1/4 B ]


2

Eq. 43

s
JOINTED

The ultimate base resistance of a rock-socketed drilled


shaft, Qb, is the product of the limiting normal stress, or bearing capacity, qult, at the base and the cross-sectional area of
the shaft base (Ab):

Bearing
Capacity
Equation No.

N/A

S << B

(g) General shear failure


with irregular failure
surface through fractured
rock mass; two or more
closely spaced joint sets

Eq. 57

FIGURE 28 Bearing capacity failure modes in rock (after Rock


Foundations 1994).

is either intact or tightly jointed (no compressible or gougefilled seams) and there are no solution cavities or voids below
the base of the pier. ONeill and Reese (1999) recommend
limiting base resistance to 2qu if the embedment into rock is
less than one diameter. In rock with high compressive
strength, the designer also must determine the structural capacity of the shaft, which may govern the allowable normal
stress at the base.
Jointed Rock Mass
When discontinuities are vertical or nearly vertical ( > 70),
and open joints are present with a spacing less than the socket
diameter (S < B, Figure 28, mode c), failure can occur (theoretically) by unconfined compression of the poorly constrained columns (Sowers 1979). Bearing capacity can be
estimated from
qult = qu = 2c tan (45 + 12 )

(44)

(43)

Other conditions that must be verified are that the rock to


a depth of at least one diameter below the base of the socket

where qu = uniaxial compressive strength and c and are


MohrCoulomb strength properties of the rock mass. If the
nearly vertical joints are closed (Figure 28, mode d), a gen-

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

44

eral wedge failure mode may develop and the bearing capacity can be approximated using Bells solution for plane
strain conditions:
qult = cN c sc +

B
N s + DN q sq
2

(45)

in which B = socket diameter; = effective unit weight of the


rock mass; D = foundation depth; Nc, N, and Nq are bearing
capacity factors; and sc, s, and sq are shape factors to account
for the circular cross section. The bearing capacity factors
and shape factors are given by:
Nc = 2 N
N = N
Nq = N

( N + 1)
( N 2 1)

(46)
(47)

(48)

N = tan 2 45 +

(49)

Nq
sc = 1 +
Nc

(50)

s = 0.6

(51)

sq = 1 + tan

(52)

In these equations (4652), the values of c and are RMS


properties, which may be difficult to determine accurately for
rock mass beneath the base of drilled shafts.
If joint spacing S is greater than the socket diameter
(Figure 28, mode e), failure occurs by splitting, leading
eventually to general shear failure. This problem has been
evaluated by Bishnoi (1968) and developed further by
Kulhawy and Goodman (1980). The solution can be
expressed by:
qult = J c Ncr

(53)

in which J = a correction factor that depends on the ratio of


horizontal discontinuity spacing to socket diameter (H/B) as
shown in Figure 29, c = rock mass cohesion, and Ncr = a bearing capacity factor given by
N cr =

2 N 2
1 + N

( cot )

S
1
N ( cot ) + 2 N
1
B
N

(54)

where N is given by Eq. 49. If the actual RMS properties are


not evaluated, Kulhawy and Carter (1992a) suggest that rock
mass cohesion in Eq. 53 can be approximated as 0.1qu, where
qu = uniaxial compressive strength of intact rock. Rock mass
cohesion can also be estimated from the HoekBrown
strength properties using Eq. 17 given in chapter two.
For the case of moderately dipping joint sets (Figure 28,
mode f, 20 < < 70), the failure surface is likely to de-

FIGURE 29 Correction factor for discontinuity spacing


(Kulhawy and Carter 1992a).

velop along the discontinuity planes. Eq. 45 can be used,


but with strength parameters representative of the joints.
Rock Foundations (1994) recommends neglecting the first
term in Eq. 45 based on the assumption that the cohesion
component of strength along the joint surfaces is highly
uncertain.
Layered Rocks
Sedimentary rock formations often consist of alternating
hard and soft layers. For example, soft layers of shale interbedded with hard layers of sandstone. Assuming the base
of the shaft is bearing on the more rigid layer, which is underlain by a soft layer, failure can occur either by flexure if
the rigid layer is relatively thick or by punching shear if the
rigid layer is thin (Figure 28, modes g and h). Both modes are
controlled fundamentally by the tensile strength of the intact
rock, which can be approximated as being on the order of
5% to 10% of the uniaxial compressive strength. According
to Sowers (1979) neither case has been studied adequately
and no analytical solution proposed. The failure modes depicted in Figure 28 merely suggest possible methods for
analysis.
Fractured Rock Mass
A rational approach for calculating ultimate bearing capacity of rock masses that include significant discontinuities
(Figure 28, mode g) is to apply Eq. 45 with appropriate
RMS properties, c' and '. However, determination of c' and
' for highly fractured rock mass is not straightforward
because the failure envelope is nonlinear and there is no stan-

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

45

dard test method for direct measurement. One possible approach is to employ the HoekBrown strength criterion
described in chapter two. The criterion is attractive because
(1) it captures the nonlinearity in the strength envelope that
is observed in jointed rock masses and (2) the required parameters can be estimated empirically using correlations to
GSI and RMR, also described in chapter two. To use this
approach, it is necessary to relate the HoekBrown strength
parameters (mb, s, and a) to MohrCoulomb strength parameters (c' and '); for example, using Eqs. 16 and 17 in
chapter two.
Alternatively, several authors (Carter and Kulhawy 1988;
Wyllie 1999) have shown that a conservative, lower-bound
estimate of bearing capacity can be made directly in terms
of HoekBrown strength parameters by assuming a failure
mode approximated by active and passive wedges; that is,
the Bell solution for plane strain. The failure mass beneath
the foundation is idealized as consisting of two zones, as
shown in Figure 30. The active zone (Zone 1) is subjected
to a major principal stress (1') coinciding at failure with the
ultimate bearing capacity (qult) and a minor principal stress
(3') that satisfies equilibrium with the horizontal stress in
the adjacent passive failure zone (Zone 2). In Zone 2, the
minor principal stress is vertical and conservatively assumed to be zero, whereas the major principal stress, acting
in the horizontal direction, is the ultimate strength according to the HoekBrown criterion. From chapter two, the
strength criterion is given by
'

'1 = '3 + qu mb 3 + s

qu

(55)

where 1' and 3' = major and minor principal effective


stresses, respectively; qu = uniaxial compressive strength of
intact rock; and mb, s, and a are empirically determined
strength parameters for the rock mass. For Zone 2, setting the
vertical stress 3' = 0 and solving Eq. 55 for 1' yields
'1 = ' H = qu s a

(56)

where H' = horizontal stress in Zone 2. To satisfy equilibrium,


the horizontal stress given by Eq. 56 is set equal to 3' in
Zone 1. Substituting 3' = qu s0.5 into Eq. 55 and considering that
1' = qult yields
a
qult = qu s a + ( mb s a + s )

(57)

The assumption of zero vertical stress at the bearing elevation may be overly conservative for many rock sockets. A
similar derivation can be carried out with the overburden
stress taken into account, resulting in the following. Let
( 'v ,b )
A = 'v ,b + qu mb
+ s
qu

(58)

where 'v,b = vertical effective stress at the socket bearing


elevation, which is also the minor principal stress in Zone 2.
Then
A

qult = A + qu mb + s

(59)

A limitation of Eqs. 5759 is that they are based on the


assumption of plane strain conditions, corresponding to
a strip footing. Kulhawy and Carter (1992a) noted that for
a circular foundation the horizontal stress between the two
assumed failure zones may be greater than for the plane
strain case, resulting in higher bearing capacity. The analysis is therefore conservative for the case of drilled shafts.
Eqs. 5759 require determination of a single rock strength
property (qu) along with an approximation of the HoekBrown
strength parameters. In chapter two, the HoekBrown strength
parameters are correlated to GSI by Eqs. 1215. This allows a
correlation to be made between the GSI of a rock mass; the
value of the coefficient mi for intact rock as given in Table 11
(chapter two), and the bearing capacity ratio qult/qu by Eq. 57.
The resulting relationship is shown graphically in Figure 31.
The bearing capacity ratio is limited by an upper-bound value
of 2.5, corresponding to the recommendation of Rowe and
Armitage (Eq. 43).

Qult

maximum qult/qu = 2.5


2.5

m i = 33
25

qusa

qult/qu

2.0

qult

20
15
10

1.5

1.0

0.5

Zone 2
(passive wedge)

Zone 1
(active
wedge)

FIGURE 30 Bearing capacity analysis.

Zone 2
(passive wedge)

0.0
10

20

30

40

50

60

70

Geological Strength Index (GSI)

FIGURE 31 Bearing capacity ratio versus GSI.

Copyright National Academy of Sciences. All rights reserved.

80

90

100

Rock-Socketed Shafts for Highway Structure Foundations

Alternatively, the bearing capacity ratio can be related approximately to the rock mass description based on RMR
(Table 9) using an earlier correlation given by Hoek and
Brown (1988). The resulting HoekBrown strength parameters (m, a, and s) are substituted into Eq. 57 to obtain the bearing capacity ratio as a function of RMR. This relationship
is shown graphically in Figure 32. Both figures are for the
case of zero overburden stress at the bearing elevation. To account for the depth of embedment and resulting surcharge
stress, Eqs. 58 and 59 can be used.

Ultimate Unit Base Resistance,


qult (MPa)

46
1000

qmax = 2.5
100

qult = a (qu)0.5
10

a = 6.6
4.8
3.0
1
0.1

Method Based on Field Load Tests

10

100

Uniaxial Compressive Strength qu (MPa)

Zhang and Einstein (1998) compiled and analyzed a database


of 39 load tests to derive an empirical relationship between
ultimate unit base resistance (qult) and uniaxial compressive
strength of intact rock (qu). Reported values of uniaxial compressive strength ranged from 0.52 MPa to 55 MPa, although
most were in the range of relatively low strength. The authors
relied on the interpretation methods of the original references
to determine ultimate base capacity and acknowledge that
some uncertainties and variabilities are likely to be incorporated into the database as a result. The results are shown
on a loglog plot in Figure 33. The linear relationship
recommended by Rowe and Armitage (1987a) is shown for
comparison. Based on statistical analysis of the data, the following recommendations are given by the authors:
Lower bound: qult = 3.0 qu

(60)

Upper bound: qult = 6.6 qu

(61)

qult = 4.8 qu

(62)

Mean:

Eqs. 6062 provide a reasonably good fit to the available


data and can be used for estimating ultimate base resistance,
3.0

FIGURE 33 Unit base resistance versus intact rock strength


(derived from Zhang and Einstein 1998).

with due consideration of the limitations associated with predicting a rock mass behavior on the basis of a single strength
parameter for intact rock. Rock mass discontinuities are not
accounted for explicitly, yet they clearly must affect bearing
capacity. By taking this empirical approach, however, rock
mass behavior is accounted for implicitly because the load
tests on which the method is based were affected by the characteristics of the rock masses. Additional limitations to the
approach given by Zhang and Einstein are noted in a discussion of their paper by Kulhawy and Prakoso (1999).
None of the analytical bearing capacity models described
above by Eqs. 44 through 59 and depicted in Figure 28 have
been evaluated and verified against results of full-scale field
load tests on rock-socketed drilled shafts. The primary reason for this is a lack of load test data accompanied by sufficient information on rock mass properties needed to apply
the models.
Canadian Geotechnical Society Method

maximum qult/qu = 2.5


2.5

qult/qu

2.0

E
D
C
B
A

1.5

1.0

0.5

A.
B.
C.
D.
E.

INTACT

VERY GOOD

GOOD

FAIR

POOR

VERY POOR

0.0

Carbonate rocks with well-developed cleavage: dolomite, limestone, marble.


Lithified argillaceous rocks: mudstone, siltstone, shale, and slate.
Arenaceous rocks with strong crystals and poorly developed crystal cleavage: sandstone and
quartzite.
Fine-grained polyminerallic igneous crystalline rocks: andesite, dolerite, diabase, and rhyolite.
Coarse-grained polyminerallic igneous and metamorphic crystalline rocks: amphibolite, gabbro,
gneiss, quartz, norite, quartz-diorite.

FIGURE 32 Bearing capacity ratio as a function of rock type


and RMR classification (see Table 9).

The Canadian Foundation Engineering Manual [Canadian


Geotechnical Society (CGS) 1985] presents a method to estimate ultimate unit base resistance of piles or shafts bearing
on rock. The CGS method is described as being applicable to
sedimentary rocks with primarily horizontal discontinuities,
where discontinuity spacing is at least 0.3 m (1 ft) and discontinuity aperture does not exceed 6 mm (0.25 in.). The
method is given by the following:
qult = 3qu K sp d

(63)

in which
K sp =

3+

sv
B

10 1 + 300

td
sv

Copyright National Academy of Sciences. All rights reserved.

(64)

Rock-Socketed Shafts for Highway Structure Foundations

47

d = 1 + 0.4

Ls
B

Qc

(65)

where
sv = vertical spacing between discontinuities,
td = aperture (thickness) of discontinuities,
B = socket diameter, and
Ls = depth of socket (rock) embedment.
A method to calculate ultimate unit base resistance from
PMT is also given by CGS as follows:
qult = K b ( p1 po ) + v

0
0.8

1
2.8

Er,r

Eb, b

(66)

where
p1 = limit pressure determined from PMT tests averaged
over a depth of two diameters above and below
socket base elevation,
po = at-rest total horizontal stress measured at base
elevation,
v = total vertical stress at base elevation; and
Kb = socket depth factor given as follows:
H/D
Kb

Ec,
c

2
3.6

3
4.2

5
4.9

7
5.2

The two CGS methods described earlier are adopted in the


draft 20006 Interim AASHTO LRFD Bridge Design Specifications (2006).
AXIAL LOAD-DISPLACEMENT BEHAVIOR

Analysis of the load-displacement behavior of a drilled shaft


is an essential step in a rational design. Design of most sockets is governed by the requirement to limit settlement to a
specified allowable value. The problem of predicting vertical displacement at the top of a rock socket has been studied
through theoretical and numerical analyses along with limited results from full-scale field load testing. Methods that
appear to have the most application to design of highway
bridge foundations are summarized in this section.
The basic problem is depicted in Figure 34 and involves
predicting the relationship between an axial compression
load (Qc) applied to the top of a socketed shaft and the
resulting axial displacement at the top of the socket (wc). The
concrete shaft is modeled as an elastic cylindrical inclusion
embedded within an elastic rock mass. The cylinder of depth
L and diameter B has Youngs modulus Ec and Poissons
ratio vc. The rock mass surrounding the cylinder is homogeneous with Youngs modulus Er and Poissons ratio vc,
whereas the rock mass beneath the base of the shaft has
Youngs modulus Eb and Poissons ratio vb. (Note: some
authors use Er to denote modulus of rock in elasticity solutions; elsewhere in this report, Er denotes modulus of intact
rock and EM is the rock mass modulus of deformation.) The

FIGURE 34 Axially loaded rock socket, elastic analysis.

shaft is subjected to a vertical compressive force Qc assumed to be uniformly distributed over the cross-sectional
area of the shaft resulting in an average axial stress b =
4Q/(B2).
Early solutions to the problem of a single compressible
pile in an elastic continuum were used primarily to study
the response of deep foundations in soil (e.g., Mattes and
Poulos 1969; Butterfield and Banerjee 1971; Randolph and
Wroth 1978). In most cases, the solutions were not directly
applicable to rock sockets because they did not cover the
typical ranges of modulus ratio (Ec/Er) or embedment ratio
(L/B) of rock sockets, but they did provide the basic
methodology for analysis of the problem. Osterberg and
Gill (1973) used an elastic finite-element formulation to analyze rock sockets with D/B ranging from zero to 4 and the
modulus ratios ranging from 0.25 to 4. Their analysis also
considered differences between the modulus of the rock beneath the base (Eb) and that along the shaft (Er). Results
showed the influence of these parameters on load transfer,
in particular the relative portion of load carried in side resistance and transmitted to the base, but did not provide a
method for predicting load-displacement behavior for design. Pells and Turner (1979) and Donald et al. (1980) conducted finite-element analyses assuming elastic and elastoplastic behaviors. Their numerical results were used to
determine values of the dimensionless influence factor (I)
that can be used to predict elastic deformation using the
general equation
wc =

Qc
I
Er B

Copyright National Academy of Sciences. All rights reserved.

(67)

Rock-Socketed Shafts for Highway Structure Foundations

48

Values of the influence factor were presented in the form


of charts for a range of modulus and embedment ratios common for rock sockets. Graphs are also provided showing
the ratio of applied load transferred to the base (Qb/Qc). These
studies provided the first practical methods for predicting the
load-displacement response of rock sockets. Their principal
limitation lies in the assumption of a full bond between the
shaft and the rock; that is, no slip. Observations from load
tests; for example, Horvath et al. (1983), show that peak side
resistance may be reached at displacements on the order of
5 mm. Rupture of the interface bond begins at this point, resulting in relative displacement (slip) between the shaft and
surrounding rock. Under service load conditions, most rock
sockets will undergo displacements that reach or exceed the
full slip condition and should be designed accordingly.
Analyses that account for both fully bonded conditions and
full slip conditions provide a more realistic model of loaddisplacement response.
Rowe and Pells (1980) conducted a theoretical study
based on finite-element analyses of rock-socketed shafts that
accounts for the possibility of slip at the shaftrock interface.
The analysis treats the shaft and rock as elastic materials, but
provides for plastic failure within the rock or the concrete
shaft and for slip at the cohesive-frictional and dilatant rock
shaft interface. At small loads, the shaft, rock, and interface
are linearly elastic and the shaft is fully bonded to the rock.
Slip is assumed to occur when the mobilized shear stress
reaches the interface strength, assumed to be governed by a
MohrCoulomb failure criterion:
= cpeak + n tanpeak

(68)

where cpeak = peak interface adhesion, n = interface normal


stress, and peak = peak interface friction angle. Once slip occurs, it is assumed that c and degrade linearly with relative
displacement between the two sides of the ruptured interface
from the peak values to residual values (cresidual, residual) at a
relative displacement r. Roughness of the interface is modeled in terms of a dilatancy angle and a maximum dilation,
and strain softening of the interface is considered. Modeling
of the interface in this way provides a good mechanistic representation of the load-displacement behavior of a rocksocketed shaft, as described in the beginning of this chapter
(Figure 19).
From these studies, Rowe and Armitage (1987a,b) prepared design charts that enable construction of a theoretical
load-displacement curve in terms of (1) the influence factor
I used to calculate axial displacement by Eq. 67 and (2) the
ratio of load Qb/Qt transmitted to the socket base, where Qb =
base load and Qt = total load applied to the top of the shaft.
Figure 35 is an example of the chart solution for a complete
socket. The charts offer a straightforward means of calculating load-displacement curves and have been used by practitioners for the design of bridge foundations. The Rowe and

FIGURE 35 Design chart for shaft displacement and base load


transfer, complete socket (Rowe and Armitage 1987a).

Armitage solutions represent a standard against which


approximate methods can be compared and verified.
Rowe and Armitage (1987a,b) developed design charts for
two contact conditions at the base of the socket: (1) a complete socket, for which full contact is assumed between the
base of the concrete shaft and the underlying rock; and (2) a
shear socket, for which a void is assumed to exist beneath the
base. These conditions are intended to model the socket
arrangements and methods of loading used in field load testing. When clean base conditions during construction can be
verified and instrumentation is provided for measuring base
load, a complete socket is assumed. Frequently, however, base
resistance is eliminated by casting the socket above the base of
the drilled socket, in which case the test shaft is modeled as a
shear socket. The boundary condition at the base of a shear
socket under axial compression is one of zero stress. The charts
given in Rowe and Armitage (1987a,b) provide a rigorous
method for analyzing rock-socket load-displacement behavior.
Closed Form Solutions

An approximate method given by Kulhawy and Carter


(1992b) provides simple, closed-form expressions that are attractive for design purposes and yield results that compare
well with those of Rowe and Armitage. For axial compression loading, the two cases of complete socket and shear
socket are treated. Solutions were derived for two portions of
the load-displacement curve depicted in Figure 19; the initial
linear elastic response (OA) and the full slip condition (re-

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

49

gion beyond point B). The closed-form expressions cannot


predict the load-displacement response between the occurrence of first slip and full slip of the shaft (AB). However, the
nonlinear finite-element results indicate that the progression
of slip along the socket takes place over a relatively small
interval of displacement. Comparisons of the bilinear curve
given by the closed-form expressions with results of Rowe
and Armitage (1987b) indicated that this simplification is
reasonably accurate for the range of rock-socket conditions
encountered in practice.
The closed-form expressions for approximating the loaddisplacement curves for complete and shear socket are given
here. For a full description of the assumptions and derivations the reader is referred to Carter and Kulhawy (1988) and
Kulhawy and Carter (1992b).
1. For the linearly elastic portion of the load-displacement curve.
(a) Shear socket (zero stress at the base):
Er Bwc 1 2 Er cosh [L ]
=
B Ec siinh [L ]
2Qc

(69)

in which wc = downward vertical displacement at


the butt (top) of the shaft and where is defined
by:
2 2L
( L )2 =
B

(70)

= ln [5(1 vr)L/B]

(71)

= Ec/Gr

(72)
(73)

Gr = Er / [2(1 + vr]

Q
wc = F1 c F2 B
Er B

(78)

in which F1 = a1(2BC2 1BC1) 4a3

(79)

c
F2 = a2
Er

(80)

C1,2 = exp[2,1L]/(exp[2L] exp[1L])


1,2 =

( 2 + 4 )
2

(b) Complete socket:

(82)
(83)

E
= a3 c B
Er

(84)

a1 = (1 + vr) + a2

(85)

1
E

a2 = (1 c ) r + (1 + r )

Ec

2 tan tan
c Er
a3 =
2 tan Ec

(86)
(87)

(b) Complete socket:


Q
wc = F3 c F4 B
Er B

(88)

in which
F3 = a1(1BC3 2BC4) 4a3

C 3 ,4 =

(74)

(81)

E B2
= a1 c
Er 4

2
c

F4 = 1 a1 1
B a2

D
Er

4
3

where Gr = elastic shear modulus of rock mass.

4 1 2 L tanh [L ]
1+
1 b B L
Gr Bwc
=
L tanh [L ]
2Qc
4 1 2 2L
+

1 b B L

2. For the full slip portion of the load-displacement curve.


(a) Shear socket:

D3,4
D4 D3

D3,4 = (1 2b ) r + 4 a3 + a1 2,1 B exp [ 2,1 L ]


Eb

(89)
(90)
(91)
(92)

The magnitude of load transferred to the base of


the shaft (Qb) is given by

where
= Gr /Gb
Gb = Eb/ [2(1 + vb]

(75)
(76)

The magnitude of load transferred to the base of


the shaft (Qb) is given by
1

4 1

Qb
1 b cosh [L ]
=
Qc 4 1 2 2 L tanh [L ]
+

1 b B L

Qb
B 2 c
= P3 + P4
Qc
Qc

(93)

in which
P3 = a1(1 2) B exp[(1 + 2)L]/(D4 D3) (94)
P4 = a2(exp[2D] exp[1L])/(D4 D3)

(77)

(95)

The solutions given previously (Eqs. 6995) are easily


implemented by spreadsheet, thus providing designers with

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

50

a simple analytical tool for assessing the likely ranges of behavior for trial designs. A spreadsheet solution provides the
opportunity to easily evaluate the effects of various input parameters on load-displacement response. When combined
with appropriate judgment and experience, this approach
represents a reasonable analysis of rock-socketed drilled
shafts. The method of Carter and Kulhawy presented herein
is also adopted in the FHWA Drilled Shaft Manual (ONeill
and Reese 1999) for analysis of load-displacement response
of single drilled shafts in rock (see Appendix C of the Manual). Reese and ONeill also present methods for predicting
load-displacement response of shafts in cohesive IGMs and
cohesionless IGMs. The equations are not reproduced here,
but are given as closed-form expressions that can be implemented easily using a spreadsheet.

Other Methods

A computer program that models the axial load-displacement


behavior of a rock-socketed shaft, based on the methods described by Seidel and Haberfield (1994) and described
briefly earlier, has been developed. The program ROCKET
requires the following input parameters:

Drained shear strength parameters of the intact rock,


Rock mass modulus and Poissons ratio,
Foundation diameter,
Initial normal stress, and
Mean socket asperity length and mean socket asperity
angle.

Although some of these parameters are determined on a


routine basis for the design of rock-socketed foundations,
asperity characteristics are not typically evaluated. Drained
triaxial tests are also not considered routine by most transportation agencies in the United States. However, this approach is promising because it provides a theoretical basis
for predicting rock-socket behavior that encompasses more
of the important parameters than the empirical approaches
now available to predict side resistance.

Combining Side and Base Resistances

A fundamental aspect of drilled shaft response to axial compression loading is that side and base resistances are mobilized
at different downward displacements. Side resistance typically reaches a maximum at relatively small displacement,
in the range of 5 to 10 mm. Beyond this level, side resistance may remain constant or decrease, depending on the
stressstrain properties of the shaft-rock interface. Ductile
behavior describes side resistance that remains constant or
decreases slightly with increasing displacement. If the interface is brittle, side resistance may decrease rapidly and
significantly with further downward displacement. One
question facing the designer is how much side resistance is

mobilized at the strength limit state. As stated by ONeill and


Reese (1999), the issue of whether ultimate side resistance
should be added directly to the ultimate base resistance to obtain an ultimate value of resistance is a matter of engineering
judgment. Responses to Question 20 of the survey (Appendix A) show a wide range in the way that side and base
resistances are combined for design of rock sockets. Several
states indicated that they follow the guidelines given by
ONeill and Reese in the FHWA Drilled Shaft Manual
(1999). Three possible approaches are described here. The
first applies to the case where load testing or laboratory shear
strength tests prove that the rock is ductile. In this case, the
ultimate values of side and base resistance are added directly.
If no field or lab tests are conducted, a fully reduced frictional shearing resistance at the interface is used to compute
side resistance and this value is added to the ultimate base resistance. The fully reduced strength is taken as the residual
shear strength of the rock = 'htanrc, where 'h = horizontal
effective stress normal to the interface and rc = residual
angle of interface friction between rock and concrete. A
value of 25 degrees is recommended in the absence of measurements and 'h is assumed to be equal to the vertical
effective stress 'v. A second approach is recommended for
cases in which base resistance is neglected in design. In this
case, the ultimate side resistance is recommended for design
at the strength limit state, unless progressive side shear failure could occur, in which case the resistance should be
reduced according to the judgment of the geotechnical
engineer.
Several states indicate in their response to Question 20
that load testing, especially using the Osterberg load cell, is
one of the ways in which the issue is addressed. Load tests
that provide independent measurements of side and base
resistance as a function of displacement and that are carried
to large displacements provide the best available data for establishing resistance values. When testing is not conducted,
the major uncertainty (i.e., judgment) is associated with the
question of whether or not side shear behavior will exhibit
significant strain softening. Research is needed to provide
guidance on what conditions are likely to produce a brittle
response of the side resistance. Most load test data in which
side resistance is measured independently do not exhibit a severe decrease in side resistance with large displacement. A
study with the objective of identifying the factors that control stressstrain behavior at the shaftrock interface at large
displacement is needed. A large amount of data have been
produced in recent years from load tests using the Osterberg
cell (O-cell) and such data would be useful for a study of
post-peak interface behavior. The results would be most
useful to practitioners if guidance could be provided on specific rock types, drilling methods, construction practices,
and ranges of confining stress that are likely to cause strain
softening at the interface. These factors could then be used
as indicators of cases for which further field or laboratory
testing is warranted.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

51

A promising technique for improving base loaddisplacement response of drilled shafts involves post-grouting
at the base (base grouting). The technique involves casting
drilled shafts with a grout delivery system incorporated into
the reinforcing cage capable of placing high pressure grout
at the base of the shaft, after the shaft concrete has cured. The
effect is to compress debris left by the drilling process, thus
facilitating mobilization of base resistance within service or
displacement limits. According to Mullins et al. (2006), base
grouting is used widely internationally, but its use in North
America has been limited. An additional potential advantage
is that the grouting procedure allows a proof test to be conducted on the shaft. Base grouting warrants further consideration as both a quality construction technique and a testing
tool for rock-socketed shafts.
For evaluation of service limit states, both side and base
resistances should be included in the analysis. Analytical
methods that can provide reasonable predictions of axial
load-deformation response, for example the Carter and Kulhawy method described in this chapter or similar methods
given by ONeill and Reese (1999), provide practical tools
for this type of analysis. All of these methods require evaluation of rock mass modulus.
CURRENT AASHTO PRACTICE

The draft 2006 Interim AASHTO LRFD Bridge Design


Specifications recommends specific methods and associated
resistance factors for evaluating side and base resistance of
rock-socketed shafts under axial load. These are summarized
in Table 20. The resistance factors are based on a calibration
study conducted by Paikowsky et al. (2004a) and additional

recommendations given by Allen (2005). Rock mass properties used with LRFD resistance factors should be based on
average values, not minimum values.
Three methods are cited for predicting ultimate unit side
resistance in rock. The first is identified as Horvath and Kenney (1979). However, the equation given in the AASHTO
Specifications is actually the original Horvath and Kenney
recommendation (Eq. 25), but with unit side resistance modified to account for RQD. A reduction factor, , is applied,
as determined by Table 13 and Table 18 of this report. This
approach was recommended by ONeill and Reese (1999).
The second method is identified as Carter and Kulhawy
(1988). The draft 2006 Interim AASHTO LRFD Bridge Design Specifications does not state explicitly the equation to
be used in connection with the Carter and Kulhawy method.
However, in the calibration study by Paikowsky et al. (2004a)
the expression used for all evaluations attributed to Carter
and Kulhawy is
fsu = 0.15 qu

(96)

in which qu = uniaxial compressive strength of rock. In their


original work, Carter and Kulhawy (1988) proposed the use
of 0.15 qu as a design check, whereas the AASHTO Specifications treat it as a design recommendation. This unintended
usage is inappropriate and does not adequately represent the
most up-to-date research based on regression analysis of the
available data on socket-side resistance. The third method
given by AASHTO is ONeill and Reese (1999). It is not
clear how this differs from the Horvath and Kenney (1979)
method because the equations given by AASHTO are all
taken directly from ONeill and Reese (1999). The equations

TABLE 20
SUMMARY OF CURRENT AASHTO METHODS AND RESISTANCE FACTORS

Nominal Axial Compressive


Resistance of Single-Drilled
Shafts

Method/Condition
Side resistance in
rock

Tip resistance in rock

1. Horvath and Kenney (1979)


2. Carter and Kulhawy (1988)
3. ONeill and Reese (1999)
1. Canadian Geotechnical Society
(1985)
2. PMT Method (Canadian
Geotechnical Society 1985)
3. ONeill and Reese (1999)

0.50

Side resistance, IGMs

1. ONeill and Reese (1999)

0.60

Tip resistance, IGMs

1. ONeill and Reese (1999)

0.55

Static load test


Compression, all
materials
Nominal Uplift Resistance of
Single-Drilled Shaft
*

Resistance
Factor
0.55
0.50
0.55

Rock

0.50
0.50

<0.70*

Horvath and Kenney (1979)


Carter and Kulhawy (1988)
Load Test

Depends on the number of load tests and site variability.

AASHTO LRFD Bridge Design Specifications, 2006 Interim.

Copyright National Academy of Sciences. All rights reserved.

0.40
0.40
0.60

Rock-Socketed Shafts for Highway Structure Foundations

52

presented in the draft 2006 Interim AASHTO LRFD Bridge


Design Specifications are not the same as those originally
proposed by Horvath and Kenney (1979) and by Carter and
Kulhawy (1988), but are nonetheless attributed to those studies. Furthermore, both studies have been superseded by more
recent research. In future calibration studies for LRFD
applications and for updates of AASHTO specifications,
consideration of alternative design equations for side resistance should be considered and the most up-to-date research
should be referenced.
The draft 2006 Interim AASHTO LRFD Bridge Design
Specifications allow the use of methods other than those
given in Table 20, especially if the method is locally recognized and considered suitable for regional conditions . . .
if resistance factors are developed in a manner that is consistent with the development of the resistance factors for the
method(s) provided in these Specifications.
AASHTO specifies resistance factors for base resistance
based on the two methods given by the Canadian Geotechnical Society (Canadian Foundation Engineering Manual
1985). The first is according to Eqs. 6365 and is a straightforward method to apply, provided the rock satisfies the criteria of being horizontally jointed and the appropriate parameters can be determined. Standard logging procedures for
rock core would normally provide the required information.
The second method is based on PMT and is given by Eq. 66.
As noted in chapter two, only a few states reported using the
PMT in rock. The third method for base resistance is ONeill
and Reese (1999) and the two equations given by AASHTO
correspond to Eq. 43 of this report for massive rock and Eq.
57 of this report for highly fractured rock.
AASHTO also allows higher resistance factors on both
side and base resistances when they are determined from a
field load test. The cost benefits achieved by using a load test
as the basis for design can help to offset the costs of conducting load tests. This issue is considered further in chapter
five. Finally, AASHTO recommends that all of the resistance
factors given in Table 20 be reduced by 20% when used for
the design of nonredundant shafts; for example, a single shaft
supporting a bridge pier.

SUMMARY

The principal factors controlling the behavior of rock-socketed


foundations under axial loading are identified and discussed.
It is concluded from this study that sufficient tools are currently available for transportation agency personnel to design
rock-socketed shafts for axial loading conditions that provide
adequate load carrying capacity without being overly conservative.
The principal performance design criteria for axial loading are (1) adequate capacity and (2) ability to limit vertical

deformation. Research published over the past 25 years has


resulted in methods for predicting ultimate side resistance of
shafts in rock that can be selected by a designer on the basis
of commonly measured geomaterial properties and that account for levels of uncertainty associated with the project.
For example, Eq. 30 with C taken equal to 1.0 provides a conservative estimate of side resistance for preliminary design
or for final design of small structures or at sites where no
additional testing is planned. If laboratory CNS testing is
conducted to measure rockconcrete interface strength,
higher values of side resistance can be justified for design. If
field load tests are conducted, they normally result in higher
side resistance values than given by Eq. 30 (with C = 1.0) and
higher resistance factors are allowed by AASHTO for results
based on load tests. If field load testing demonstrates that a
particular construction technique; for example, artificial
roughening the walls of a socket, can increase side resistance,
then it may be possible to justify the use of Eq. 30 with values of C higher than 1.0.
Rational methods are available for estimating ultimate
base resistance of rock sockets. A first order approximation
based on strength of intact rock is given by Zhang and
Einstein (1998) (see Figure 33). For fractured rock, a reasonable estimate can be made if the GSI (or RMR) is evaluated (see Figures 31 and 32). Although most states surveyed
do not currently use GSI and RMR, the parameters required
for its implementation can be obtained during the course of
standard core logging procedures. The method recommended by CGS and adopted by AASHTO is applicable to
moderately jointed sedimentary rocks, which is the most
commonly encountered rock type for rock-socketed foundations. A method based on PMT provides another practical
approach for calculating base resistance.
A source of uncertainty in rock-socket design stems from
attempting to combine side and base resistances at a specific
value of downward displacement; for example, at the specified limiting value of settlement or at the strength limit state.
A relatively straightforward analysis based on elastic continuum theory, as given by Carter and Kulhawy (1988), is presented in the form of closed-form expressions that predict
axial load-deformation and base load transfer for typical
conditions encountered in practice. Similar analytical approaches for IGMs are given by ONeill and Reese (1999).
These equations are easily implemented in spreadsheet or
other convenient form and allow designers to make rational
estimates of load carried by both side and base at specified
displacements. The survey questionnaire shows that this
method is used by some state DOTs. The approach should be
evaluated further against field load test measurements and, if
verified, used more widely. Alternatively, the design charts
given in Rowe and Armitage (1987a,b) provide a rational
means of estimating axial load-displacement behavior and
base load transfer. The charts are based on rigorous numerical modeling and are the benchmark against which the Carter

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

53

and Kulhawy closed-form expressions were evaluated. However, the charts are more cumbersome to use. A computer
program that models the full load-displacement curve,
ROCKET, is available, but requires input parameters that
normally are not determined by transportation agencies, such
as triaxial strength properties and socket roughness parameters. However, for agencies interested in obtaining the
required material properties, this program offers an effective
method for axial load-deformation analysis.

Methods for calculating nominal (ultimate) unit side and


base resistances and associated resistance factors according
to the Interim 2006 AASHTO LRFD Bridge Design Specifications are summarized in Table 20. Considering the information identified by the literature review, in particular recent
studies on correlation equations for unit side resistance, a
suggested improvement in future specifications would be to
consider design methods recommended by the more recent
studies for inclusion and calibration to LRFD.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

54

CHAPTER FOUR

DESIGN FOR LATERAL LOADING

SCOPE

Twenty-two states indicated in the survey that lateral


loading considerations govern the design of drilled shafts
in rock or IGM on at least some of their projects. Several
states responded that lateral loading governs 100% of
their designs. The survey also demonstrated that the most
widely used analysis is the p-y method, although other
methods are also used. In this chapter, analytical models
identified by the literature search and by the survey are reviewed and evaluated for their applicability to rock sockets. Structural issues associated with rock-socketed shafts
are considered.

that are then used by GD to perform revised lateral load


analyses. In addition, the Bridge Office may conduct their
own analyses using soil and rock-structure interaction models with soil and rock properties provided by GD. The
Bridge Office uses the modeling results to establish design
parameters for drilled shaft reinforced concrete. According
to the 2006 Interim AASHTO LRFD Bridge Design Specifications, the strength limit state for lateral resistance of
deep foundations is structural only. The basic assumption
is that failure of the soil/rock does not occur; instead, the
geomaterials continue to deform at constant or slightly
increasing resistance. Failure occurs when the foundation
reaches the structural limit state, defined as the loading at
which the nominal combined bending and axial resistance is
reached.

DESIGN PROCESS

Deep foundations supporting bridge structures may be subjected to lateral loading from a variety of sources, including
earth pressures, centrifugal forces from moving vehicles,
braking forces, wind loading, flowing water, waves, ice, seismic forces, and impact. Reese (1984) describes numerous
examples of bridges, overhead sign structures, and retaining
structures as typical examples of transportation facilities that
must sustain significant lateral loading. Drilled shafts are often selected for such structures because they can be designed
to sustain lateral loading by proper sizing of the shaft and by
providing a sufficient amount of reinforcing steel to resist the
resulting bending moments.
Design for lateral loading of drilled shafts requires significant interaction between geotechnical and structural
engineers. As described in chapter one, the Bridge Office
(structural) is responsible for structural analysis and design
of the superstructure and the foundations. However, to model
foundation response to lateral loading it is necessary to account for soil/rock-structure interaction. The Geotechnical
Division (GD) normally conducts foundation analysis using
the models described in this chapter. For preliminary foundation design, geotechnical modeling of foundation response by the GD is used to provide the Bridge Office with
information such as depth of fixity, trial designs (diameter
and depth) of drilled shafts, and equivalent lateral spring
values for use in seismic analysis of the superstructure. The
Bridge Office conducts analyses of the superstructure based
on models that include fixed-end columns at the depth of
fixity. The structural analysis may result in revised loads

Axial loading in both compression and uplift requires structural analysis and reinforced-concrete design for drilled shafts.
When lateral loading is not significant, structural design is reasonably straightforward. When lateral loading is significant,
the combined effects of lateral and axial loading are analyzed
using models described in this chapter, which account for the
effect of axial load by treating the shaft as a beam column. For
this reason, structural design issues associated with rocksocketed shafts are addressed in this chapter.

ROCK-SOCKETED FOUNDATIONS
FOR LATERAL LOADING

Rock-socketed shafts provide significant benefits for carrying lateral loads. Embedment into rock, in most cases,
reduces the lateral displacements substantially compared
with a deep foundation in soil. To take full advantage of
rock-socketed drilled shafts, designers must have confidence in the analytical tools used for design. The survey
questionnaire shows that traditional methods of analysis
for lateral loading of piles and drilled shafts in soil are the
most widely used methods currently employed for rock
sockets. Recent research has led to some advancements for
applying these methods to rock. In addition, several researchers have proposed new analytical methods that provide
designers with useful tools for predicting load-displacement
response and/or structural response of the reinforcedconcrete shaft. Each method has advantages and disadvantages
for design purposes and these are discussed in the following
sections.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

55
ANALYTICAL METHODS

A laterally loaded deep foundation is the classic example of a


soilstructure interaction problem. The soil or rock reaction
depends on the foundation displacement, whereas displacement is dependent on the soil or rock response and flexural
rigidity of the foundation. In most methods of analysis, the
foundation is treated as an elastic beam or elastic beamcolumn. The primary difference in analytical methods used to
date lies in the approach used to model the soil and/or rock
mass response. Methods of analysis fall into two general categories: (1) subgrade reaction and (2) elastic continuum theory.
Subgrade reaction methods treat the deep foundation analytically as a beam on elastic foundation. The governing differential equation (Hetenyi 1946) is given by
EI

d4y
d2y
+ Pz 2 p w = 0
4
dz
dz

(97)

in which EI = flexural rigidity of the deep foundation, y =


lateral deflection of the foundation at a point z along its length,
Pz = axial load on the foundation, p = lateral soil/rock reaction per unit length of foundation, and w = distributed load
along the length of the shaft (if any). In the most commonly
used form of subgrade reaction method, the soil reactiondisplacement response is represented by a series of independent nonlinear springs described in terms of p-y curves
(Reese 1984). A model showing the concept is provided in
Figure 36. The soil or rock is replaced by a series of discrete
mechanisms (nonlinear springs) so that at each depth z the
soil or rock reaction p is a nonlinear function of lateral deflection y. Ideally, each p-y curve would represent the
stressstrain and strength behavior of the soil or rock, effects
of confining stress, foundation diameter and depth, position
of the water table, and any other factors that determine the

FIGURE 36 Subgrade reaction model based on p-y curves


(Reese 1997).

actual soil/rock reaction. In practice, a great deal of effort and


research has been aimed at developing methods for selecting
appropriate p-y curves. All of the proposed methods are empirical and there is no independent test to determine the relevant p-y curve. The principal limitations of the p-y method
normally cited are that: (1) theoretically, the interaction of
soil or rock between adjacent springs is not taken into account (no continuity), and (2) the p-y curves are not related
directly to any measurable material properties of the soil or
rock mass or of the foundation. Nevertheless, full-scale load
tests and theory have led to recommendations for establishing p-y curves for a variety of soil types (Reese 1984).
The method is attractive for design purposes because of the
following:
Ability to simulate nonlinear behavior of the soil or
rock;
Ability to follow the subsurface stratigraphy (layering)
closely;
Can account for the nonlinear flexural rigidity (EI) of a
reinforced-concrete shaft;
Incorporates realistic boundary conditions at the top of
the foundation;
Solution provides deflection, slope, shear, and moment
as functions of depth;
Solution provides information needed for structural design (shear and moment); and
Computer solutions are readily available.
Boundary conditions that can be applied at the top of the
foundation include: (1) degree of fixity against rotation or
translation and (2) applied loads (moment, shear, and axial).
With a given set of boundary conditions and a specified
family of p-y curves, Eq. 97 is solved numerically, typically
using a finite-difference scheme. An iterative solution is
required to incorporate the nonlinear p-y curves as well as
the nonlinear moment versus EI relationship (material and
geometric nonlinearities) for reinforced-concrete shafts.
The elastic continuum approach for laterally loaded deep
foundations was developed by Poulos (1971), initially for
analysis of a single pile under lateral and moment loading at the
pile head. The numerical solution is based on the boundary
element method, with the pile modeled as a thin elastic strip
and the soil modeled as a homogeneous, isotropic elastic material. This approach was used to approximate socketed piles
by Poulos (1972) by considering two boundary conditions at
the tip of the pile: (1) the pile is completely fixed against rotation and displacement at the tip (rock surface) and (2) the
pile is free to rotate but fixed against translation (pinned) at
the tip. The fixed pile tip condition was intended to model a
socketed deep foundation, whereas the pinned tip was intended to model a pile bearing on, but not embedded into,
rock. Although these tip conditions do not adequately model
the behavior of many rock-socketed shafts, the analyses
served to demonstrate some important aspects of socketed
deep foundations. For relatively stiff foundations, which

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

56

applies to many drilled shafts, considerable reduction in displacement at the pile head can be achieved by socketing,
especially if the effect of the socket is to approximate a
fixed condition at the soil/rock interface.
The elastic continuum approach was further developed by
Randolph (1981) through use of the finite-element method
(FEM). Solutions presented by Randolph cover a wide range
of conditions for flexible piles and the results are presented in
the form of charts as well as convenient closed-form solutions
for a limited range of parameters. The solutions do not adequately cover the full range of parameters applicable to rocksocketed shafts used in practice. Extension of this approach
by Carter and Kulhawy (1992) to rigid shafts and shafts of
intermediate flexibility, as described subsequently, has led to
practical analytical tools based on the continuum approach.
Sun (1994) applied elastic continuum theory to deep foundations using variational calculus to obtain the governing differential equations of the soil and pile system, based on the
Vlasov model for a beam on elastic foundation. This approach
was extended to rock-socketed shafts by Zhang et al. (2000),
and is also described in this chapter.

p-y Method for Rock Sockets

The p-y method of analysis, as implemented in various computer codes, is the single most widely used method for
design of drilled shafts in rock. Responses to the survey questionnaire for this study showed that 28 U.S. state transportation
agencies (of 32 responding) use this method. The analytical
procedure is dependent on being able to represent the response of soil and rock by an appropriate family of p-y
curves. The only reliable way to verify p-y curves is through
instrumented full-scale load tests. The approach that forms
the basis for most of the published recommendations for p-y
curves in soil is to instrument deep foundations with strain
gages to determine the distribution of bending moment over
the length of the foundation during a load test. Assuming that
the bending moment can be determined reliably from strain
gage measurements, the moment as a function of depth can
be differentiated twice to obtain p and integrated twice to
obtain y. Measured displacements at the foundation head
provide a boundary condition at that location. The p-y curves
resulting from analysis of field load tests have then been correlated empirically to soil strength and stressstrain properties determined from laboratory and in situ tests.

have been conducted and published to date. This lack of


verification can be viewed as a limitation on use of the p-y
method for rock-socketed drilled shafts. A single study by
Reese (1997) presents the only published criteria for selection of p-y curves in rock. A few state DOTs have developed
in-house correlations for p-y curves in rock.
Reese (1997)
Reese proposed interim criteria for p-y curves used for
analysis of drilled shafts in rock. Reese cautions that the
recommendations should be considered as preliminary because of the meager amount of load test data on which they
are based. The criteria are summarized as follows. For
weak rock, defined as rock with unconfined compressive
strength between 0.5 MPa and 5 MPa, the shape of the p-y
curve, as shown in Figure 37, can be described by the
following equations. For the initial linear portion of the
curve
for y yA

p = Kiry

(98)

For the transitional, nonlinear portion


p=

pur y
2 yrm

0.25

for y yA, p pur

yrm = krmB

(99)
(100)

and when the ultimate resistance is reached


p = pur

(101)

where
Kir = initial slope of the curve,
pur = the rock mass ultimate resistance,
B = shaft diameter, and
krm is a constant ranging from 0.0005 to 0.00005 that serves
to establish the overall stiffness of the curve.

An alternative approach for deducing p-y curves from load


tests is to measure the shape of the deformed foundation; for
example, using slope inclinometer measurements and fitting
p-y curves to obtain agreement with the measured displacements. This approach is described by Brown et al. (1994).
Very few lateral load tests on drilled shafts in rock, with
the instrumentation necessary to back-calculate p-y curves,

FIGURE 37 Proposed p-y curve for weak rock (Reese 1997).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

57

The value of yA corresponding to the upper limit of the initial


linear portion of the curve is obtained by setting Eq. 98 equal
to Eq. 99, yielding

pur
yA =

0.25
2 ( yrm ) K ir

1.333

(102)

The following expression is recommended for calculating


the rock mass ultimate resistance:
x
pur = r qu B 1 + 1.4 r

B
pur = 5.2rquB

for 0 xr 3B

for xr 3B

(103)
(104)

in which qu = uniaxial compressive strength of intact rock,


r = strength reduction factor, and xr = depth below rock
surface. Selection of r is based on the assumption that fracturing will occur at the surface of the rock under small deflections, thus reducing the rock mass compressive strength. The
value of r is assumed to be one-third for RQD of 100 and to
increase linearly to unity at RQD of zero. The underlying assumption is that, if the rock mass is already highly fractured,
then no additional fracturing with accompanying strength
loss will occur. However, this approach appears to have a
fundamental shortcoming in that it relies on the compressive
strength of the intact rock and not the strength of the rock
mass. For a highly fractured rock mass (low RQD) with a
high-intact rock strength, it seems that the rock mass strength
could be overestimated.
The initial slope of the p-y curve, Kir, is related to the initial elastic modulus of the rock mass as follows:
K ir kir Eir

weak rock. The first load test was located at Islamorada,


Florida. A drilled shaft, 1.2 m in diameter and 15.2 m long,
was socketed 13.3 m into a brittle, vuggy coral limestone. A
layer of sand over the rock was retained by a steel casing and
lateral load was applied 3.51 m above the rock surface. The
following values were used in the equations for calculating
the p-y curves: qu = 3.45 MPa, r = 1.0, Eir = 7,240 MPa, krm =
0.0005, B = 1.22 m, L = 15.2 m, and EI = 3.73 106 kN-m2.
Comparison of pile head deflections measured during the
load test and from p-y analyses are shown in Figure 38. With
a constant value of EI as given above, the analytical results
show close agreement with the measured displacements up
to a lateral load of about 350 kN. By reducing the values of
flexural rigidity in portions of the shaft subject to high moments, the p-y analysis was adjusted to yield deflections that
agreed with the measured values at loads higher than 350 kN.
The value of krm = 0.0005 was also determined on the basis
of establishing agreement between the measured and predicted displacements.
The second case analyzed by Reese (1997) is a lateral load
test conducted on a drilled shaft socketed into sandstone at a
site near San Francisco. The shaft was 2.25 m in diameter
with a socket length of 13.8 m. Rock mass strength and modulus values were estimated from PMT results. Three zones of
rock were identified and average values of strength and modulus were assigned to each zone. The sandstone is described
as medium-to-fine-grained, well sorted, thinly bedded, very
intensely to moderately fractured. Twenty values of RQD
were reported, ranging from zero to 80, with an average of
45. For calculating p-y curves, the strength reduction factor
r was taken as unity, on the assumption that there was little chance of brittle fracture. Values of the other parameters

(105)

where Eir = rock mass initial elastic modulus and kir = dimensionless constant given by
400 xr

kir = 100 +

3 B
kir = 500

for xr > 3B

for 0 xr 3B

(106)
(107)

The expressions for kir were determined by fitting a p-y


analysis to the results of a field load test (back-fitting) in
which the initial rock mass modulus value was determined
from PMTs. The method recommended by Reese (1997) is
to establish Eir from the initial slope of a pressuremeter curve.
Alternatively, Reese suggests the correlation given by Bieniawski (1978) between rock mass modulus, modulus of intact
rock core, and RQD, given as expression 2 in Table 12 (chapter two) of this report. According to Reese (1997), rock mass
modulus EM determined this way is assumed to be equivalent
to Eir in Eq. 105.
Results of load tests at two sites are used by Reese (1997)
to fit p-y curves according to the criteria given previously for

FIGURE 38 Measured and analytical deflection curves for shaft


in vuggy limestone (Reese 1997).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

58

used for p-y curve development were: krm = 0.00005; qu =


1.86 MPa for depth of 03.9 m, 6.45 MPa for depth of
3.98.8 m, and 16.0 MPa for depth of more than 8.8 m; Eir =
10qu (MPa) for each layer, B = 2.25 m, and EI = 35.15 x 103
MN-m2. The value of krm was adjusted to provide agreement
between displacements given by the p-y method of analysis
and measured displacements from the load test.
Figure 39 shows a comparison of the measured loaddisplacement curve with results produced by the p-y
method of analysis, for various methods of computing the
flexural rigidity (EI) of the test shaft. Methods that account
for the nonlinear relationship between bending moment and
EI provide a better fit than p-y analysis with a constant vale
of EI. The curve labeled Analytical in Figure 39 was obtained using an analytical procedure described by Reese to
incorporate the nonlinear momentEI relationships directly
into the numerical solution of Eq. 97, whereas the curve labeled ACI incorporates recommendations by the American Concrete Institute for treating the nonlinear momentEI
behavior.

range of 0.55 MPa. The user assigns a value to krm. The documentation (Ensoft, Inc. 2004) recommends to:
. . . examine the stressstrain curve of the rock sample. Typically, the krm is taken as the strain at 50% of the maximum
strength of the core sample. Because limited experimental data
are available for weak rock during the derivation of the p-y
criteria, the krm from a particular site may not be in the range
between 0.0005 and 0.00005. For such cases, you may use the
upper bound value (0.0005) to get a larger value of yrm, which in
turn will provide a more conservative result.

Fitting of p-y curves to the results of the two load tests as


described previously forms the basis for recommendations
that have been incorporated into the most widely used computer programs being used by state DOTs for analysis of laterally loaded rock-socketed foundations. The program
COM624 (Wang and Reese 1991) and its commercial version,
LPILE (Ensoft, Inc. 2004), allow the user to assign a limited
number of soil or rock types to each subsurface layer. One of
the options is weak rock. If this geomaterial selection is
made, additional required input parameters are unit weight,
modulus, uniaxial compressive strength, RQD, and krm. The
program then generates p-y curves using Eqs. 98107. The
program documentation recommends assigning weak rock
to geomaterials with uniaxial compressive strengths in the

The criteria recommended for p-y curves in the LPILEPLUS


users manual (Ensoft Inc. 2004) for strong rock is illustrated in Figure 40. Strong rock is defined by a uniaxial
strength of intact rock qu 6.9 MPa. In Figure 40, su is
defined as one-half of qu and b is the shaft diameter. The p-y
curve is bilinear, with the break in slope occurring at a
deflection y corresponding to 0.04% of the shaft diameter.
Resistance (p) is a function of intact rock strength for both
portions of the curve. The criterion does not account explicitly for rock mass properties, which would appear to limit its
applicability to massive rock. The authors recommend verification by load testing if deflections exceed 0.04% of the
shaft diameter, which would exceed service limit state criteria in most practical situations. Brittle fracture of the rock is
assumed if the resistance p becomes greater than the shaft
diameter times one-half of the uniaxial compressive strength
of the rock. The deflection y corresponding to brittle fracture
can be determined from the diagram as 0.0024 times the shaft
diameter. This level of displacement would be exceeded in
many practical situations. It is concluded that the recommended criteria applies only for very small lateral deflections
and is not valid for jointed rock masses. Some practitioners
apply the weak rock criteria, regardless of material strength,
to avoid the limitations cited earlier. The authors state that
the p-y curve shown in Figure 40 should be employed with

FIGURE 39 Measured and analytical deflection curves, socket


in sandstone (Reese 1997).

FIGURE 40 Recommended p-y curve for strong rock (Ensoft,


Inc. 2004).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

59

caution because of the limited amount of experimental data


and because of the great variability in rock.

p ult

The survey questionnaire for this study found that 28


agencies use either COM624 or LPILEPLUS for analysis of
rock-socketed shafts under lateral loading.
The following observations are based on a review of the
literature:
Existing published criteria for p-y curves in rock are
based on a very limited number (two) of full-scale field
load tests,
Recommendations for selecting values of input parameters required by the published criteria are vague and
unsubstantiated by broad experience, and
The p-y method of analysis is being used extensively
despite these sources of uncertainty.
It is therefore concluded that research is needed and should be
undertaken with the objective of developing improved criteria
for p-y curves in rock. The research should include full-scale
field load tests on instrumented shafts, much in the same way
that earlier studies focused on the same purpose for deep foundations in soil. The p-y curve parameters should be related to
rock mass engineering properties that can be determined by
state transportation agencies using available site and material
characterization methods, as described in chapter two.

Current Research by State Agencies


The literature review and the survey identified two state transportation agencies (North Carolina and Ohio) with research in
progress aimed at improving the methodology for constructing p-y curves for weathered rock. The North Carolina study
is described in a draft report by Gabr et al. (2002) and the
Ohio study is summarized in a paper that was under review
at the time of this writing, by Liang and Yang (2006). Both
studies present recommendations for p-y curves based on a
hyperbolic function. Two parameters are required to characterize a hyperbola, the initial tangent slope and the asymptote.
For the proposed hyperbolic p-y models, these correspond to
the subgrade modulus (Kh) and the ultimate resistance (pult),
as shown in Figure 41. The hyperbolic p-y relationship is then
given as
p=

y
FIGURE 41 Hyperbolic p-y curve.

model. Tests were performed on shafts in Piedmont weathered profiles of sandstone, mica schist, and crystalline rock.
Finite-element modeling was used to calibrate a p-y curve
model incorporating subgrade modulus as determined from
PMT readings and providing close agreement with strains
and deflections measured in the load tests. The model was
then used to make forward predictions of lateral load response for subsequent load tests on socketed shafts at two
locations in weathered rock profiles different than those used
to develop the model.
The procedure for establishing values of subgrade modulus
Kh involves determination of the rock mass modulus (EM) from
PMT measurements. The coefficient of subgrade reaction is
then given by:
kh =

0.65E M E M B 4
B (1 vr2 ) Es I s

1
12

(108)

A summary of the two studies, including recommendations for selection of the required parameters (Kh and pult),
is presented.
In the North Carolina study, results of six full-scale field
load tests, at three different sites, were used to develop the

units : l 3

(109)

in which B = shaft diameter, EM = rock mass modulus, vr =


Poissons ratio of the rock, and Es and Is are modulus and moment of inertia of the shaft, respectively. A procedure is
given by Gabr et al. (2002) for establishing the point of rotation of the shaft. For p-y curves above the point of rotation,
subgrade modulus is equal to the coefficient of subgrade reaction times the shaft diameter or
Kh = kh B

y
1
y
+
K h pult

kh

(110)

For depths below the point of rotation, a stiffer lateral subgrade reaction is assigned and the reader is referred to Gabr
et al. (2002) for the equations. An alternative procedure is
presented for cases where rock mass modulus is determined
using the empirical correlation given by Hoek and Brown
(1997) and presented previously as expression 7 in Table 12
of chapter two. In that expression, rock mass modulus is correlated with GSI and uniaxial compressive strength of intact
rock (qu).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

60

The second required hyperbolic curve parameter is the


asymptote of the p-y curve, which is the ultimate resistance
pult. The proposed expression is given by
pult = (pL + max)B

(111)

where pL = limit normal stress and max = shearing resistance


along the side of the shaft. Gabr et al. adopted the following
recommendation of Zhang et al. (2000) for unit side resistance:

max = 0.20 qu ( MPa)

(112)

The limit normal stress is estimated on the basis of Hoek


Brown strength parameters as determined through correlations with RMR and GSI, and is given by
'z
pL = 'z + qu mb
+ s
qu

(113)

in which ' = effective unit weight of the rock mass, z = depth


from the rock mass surface, and the coefficients mb, s, and a
are the HoekBrown coefficients given by Eqs. 1215 in
chapter two.

ior of the concrete shaft, which reduces the predicted deflections more significantly than the p-y criteria. One of the limitations of the p-y criterion proposed by Gabr et al. (2002) is
that it is based on analyses in which EI is taken as a constant.
For proper analysis of soilrockstructure interaction during
lateral loading, the nonlinear momentEI relationship should
be modeled correctly.
The North Carolina DOT also reports using the program
LTBASE, which analyzes the lateral load-displacement
response of deep foundations as described by Gabr and
Borden (1988). The analysis is based on the p-y method, but
also accounts for base resistance by including a vertical
resistance component mobilized by shaft rotation and horizontal shear resistance, as illustrated in Figure 43. Base resistance becomes significant as the relative rigidity of the
shaft increases and as the slenderness ratio decreases. For
relatively rigid rock sockets, mobilization of vertical and
shear resistance at the tip could increase overall lateral capacity significantly, and base resistance effects should be
considered. Gabr et al. (2002) stated that the hyperbolic WR
p-y curve model is now incorporated into LTBASE, but no
results were given.

Results of one of the field load tests conducted for the purpose of evaluating the predictive capability of the proposed
weak rock (WR) model is shown in Figure 42. The analyses
were carried out using the program LPILE. Analyses were also
conducted using p-y curves as proposed by Reese (1997), described previously, as well as several other p-y curve recommendations. The proposed model based on hyperbolic p-y
curves derived from PMT measurements (labeled dilatometer
in Figure 42) shows good agreement with the test results. The
authors (Gabr et al. 2002) attributed the underpredicted displacements obtained using the Reese criteria to the large values
of the factor kir predicted by Eqs. 106 and 107. However, the
analysis did not incorporate the nonlinear momentEI behav-

In the Ohio DOT study, Liang and Yang (2006) also propose a hyperbolic p-y curve criterion. The derivation is based
on theoretical considerations and finite-element analyses.
Results of two full-scale, fully instrumented field load tests
are compared with predictions based on the proposed p-y
curve criterion. The initial slope of the hyperbolic p-y curve
is given by the following semi-empirical equation:

FIGURE 42 Measured lateral load deflection versus predicted


(Gabr et al. 2002).

FIGURE 43 Base deformation as a function of shaft rotation


(Gabr and Borden 1988).

B
K h = EM
Bref

2 v Es I s
r
E B 4
e
M

0.284

Copyright National Academy of Sciences. All rights reserved.

(114)

Rock-Socketed Shafts for Highway Structure Foundations

61

in which Bref = a reference diameter of 0.305 m (1 ft) and all


other terms are as defined above for Eq. 109. Liang and Yang
(2006) recommend modulus values EM from PMTs for use in
Eq. 114, but in the absence of PMT measurements they present the following correlation equation relating EM to modulus of intact rock and GSI:
EM =

GSI
Er 21
e .7
100

(115)

where Er = elastic modulus of intact rock obtained during


uniaxal compression testing of core samples. Eq. 115 is also
expression 8 of Table 12 in chapter two. Liang and Yang
(2006) present two equations for evaluating pult. The first corresponds to a wedge failure mode, which applies to rock
mass near the ground surface. The second applies to rock
mass at depth and is given by
2

pult = pL + max pA B
4

(116)

where pL = limit normal stress, max = shearing resistance along


the side of the shaft, and pA = horizontal active pressure.
Eq. 116 is similar to Eq. 111 (Gabr et al. 2002), but accounts
for active earth pressure acting on the shaft. Both methods incorporate the HoekBrown strength criterion for rock mass to
evaluate the limit normal stress pL, and both rely on correlations with GSI to determine the required HoekBrown
strength parameters. In the Liang and Yang (2006) approach,
pult at each depth is taken as the smaller of the two values
obtained from the wedge analysis or by Eq. 116.
A source of uncertainty in all of the proposed p-y criteria
derives from the choice of method for selecting rock mass
modulus when more than one option is available. For example, using the pressuremeter and GSI data reported by Gabr
et al. (2002) significantly different values of modulus are obtained for the same site. In some cases, the measured shaft
load-displacement response (from load testing) shows better
agreement with p-y curves developed from PMT modulus,
whereas another load test shows better agreement with p-y
curves developed from GSI-derived modulus. Proper selection of rock mass modulus for foundation design is one of
the challenges for design of rock-socketed shafts, as pointed
out in chapter two. This issue becomes most important when
p-y curves for lateral load analysis are based on rock mass
modulus. Both the Reese (1997) criteria and the hyperbolic
criteria require rock mass modulus to determine the slope of
p-y curves.
The North Carolina and Ohio programs provide examples
of state DOT efforts to advance the state of practice in design
of rock-socketed foundations. The programs incorporate
careful site investigations using available methods for characterizing rock mass engineering properties (RMR, GSI) as
well as in situ testing (PMT). Both programs are based on

analysis of full-scale field load tests on instrumented shafts.


However, the proposed equations for generating p-y curves
differ between the two proposed criteria and both models will
result in different load-displacement curves. It is not clear if
either model is applicable to rock sockets other than those
used in its development. Both sets of load tests add to the
database of documented load tests now available to researchers. A useful exercise would be to evaluate the North
Carolina proposed criteria against the Ohio load test results
and vice versa.
Florida Pier
Several states reported using other computer programs that
are based on the p-y method of analysis. Seven agencies report using the Florida Bridge Pier Analysis Program
(FBPIER) for analysis of rock-socketed shafts. Of those
seven, six also report using COM624 and/or LPILE. The
FBPIER, described by Hoit et al. (1997), is a nonlinear, finiteelement analysis program designed for analyzing bridge pier
substructures composed of pier columns and a pier cap supported on a pile cap and piles or shafts including the soil (or
rock). FBPIER was developed to provide an analytical tool
allowing the entire pier structure of a bridge to be analyzed
at one time, instead of multiple iterations between foundation
analysis programs (e.g., COM624) and structural analysis
programs. Basically, the structural elements (pier column,
cap, pile cap, and piles) are modeled using standard structural
finite-element analysis, including nonlinear capabilities
(nonlinear MEI behavior), whereas the soil response is
modeled by nonlinear springs (Figure 44). Axial soil response is modeled in terms of t-z curves, whereas lateral
response is modeled in terms of p-y curves. The program has
built-in criteria for p-y curves in soil, based on published
recommendations and essentially similar to those employed
in LPILE. User-defined p-y curves can also be specified. To
simulate rock, users currently apply the criteria for either soft
clay (Matlock 1970) or stiff clay (Reese and Welch 1975)
but with strength and stiffness properties of the rock, or userdefined curves are input. Research is underway to incorporate improved p-y curve criteria into FBPIER, specifically for

FIGURE 44 Florida pier model for structure and foundation


elements (Hoit et al. 1997).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

62

Florida limestone, as described by McVay and Niraula (2004).


Centrifuge tests were conducted in which instrumented model
shafts embedded in a synthetic rock (to simulate Florida
limestone) were subjected to lateral loading. Strain gage
measurements were used to back-calculate p-y curves,
which are presented in normalized form, with p normalized
by shaft diameter and rock compressive strength (p/Bqu) and
y normalized by shaft diameter (y/B). There is no analytical
expression recommended for new p-y curve criteria and the
report recommends that field testing be undertaken on fullsize drilled shafts to validate the derived p-y curves established from the centrifuge tests before they are employed in
practice.

Strain Wedge Model


The strain wedge (SW) model has been applied to laterally
loaded piles in soil, as described by Ashour et al. (1998). The
2006 Interim AASHTO LRFD Bridge Design Specifications
identify the SW model as an acceptable method for lateral
load analysis of deep foundations. The 3-D soilpile interaction behavior is modeled by considering the lateral resistance
that develops in front of a mobilized passive wedge of soil at
each depth. Based on the soil stressstrain and strength properties, as determined from laboratory triaxial tests, the horizontal soil strain () in the developing passive wedge in front
of the pile is related to the deflection pattern (y) versus depth.
The horizontal stress change (H) in the developing passive
wedge is related to the soilpile reaction (p), and the nonlinear soil modulus is related to the nonlinear modulus of subgrade reaction, which is the slope of the p-y curve. The SW
model can be used to develop p-y curves for soil that show
good agreement with load test results (Ashour and Norris
2000). Theoretically, the SW model overcomes some of the
limitations of strictly empirically derived p-y curves because
the soil reaction (p) at any given depth depends on the response of the neighboring soil layers (continuity) and properties of the pile (shape, stiffness, and head conditions).
Ashour et al. (2001) proposed new criteria for p-y curves in
weathered rock for use with the SW model. The criteria are
described by the authors as being based on the weak rock
criteria of Reese (1997) as given by Eqs. 98-104, but modified to account for the nonlinear rock mass modulus and the
strength of the rock mass in terms of MohrCoulomb
strength parameters c and . Ashour et al. (2001) reported
good agreement between the SW analysis and a field load
test reported by Brown (1994). One state DOT (Washington)
reports using the computer program (S-Shaft) based on the
SW model that incorporates the p-y curve criteria for weathered rock. However, the program has not yet been used for
design of a socketed shaft (J. Cuthbertson, personal communication, Sep. 30, 2005). The SW model and proposed
p-y criteria of Ashour et al. (2001) warrant further consideration and should be evaluated against additional field load
test results (e.g., the tests reported by Gabr et al. 2002 and
Liang and Yang 2006).

Continuum Models for Laterally Loaded Sockets

Carter and Kulhawy (1992)


Carter and Kulhawy (1988, 1992) studied the behavior of flexible and rigid shafts socketed into rock and subjected to lateral
loads and moments. Solutions for the load-displacement relations were first generated using finite-element analyses. The
finite-element analyses followed the approach of Randolph
(1981) for flexible piles under lateral loading. Based on the
FEM solutions, approximate closed-form equations were
developed to describe the response for a range of rock-socket
parameters typically encountered in practice. The results provide a first-order approximation of horizontal groundline displacements and rotations and can incorporate an overlying soil
layer. The method is summarized as follows.
Initially, consider the case where the top of the shaft corresponds to the top of the rock layer (Figure 45). The shaft is
idealized as a cylindrical elastic inclusion with an effective
Youngs modulus (Ee), Poissons ratio (vc), depth (D), and diameter (B), subjected to a known lateral force (H), and an
overturning moment (M). For a reinforced-concrete shaft
having an actual flexural rigidity equal to (EI)c, the effective
Youngs modulus is given by
Ee =

( EI )c
B 4
64

(117)

It is assumed that the elastic shaft is embedded in a homogeneous, isotropic elastic rock mass, with properties Er
and vr. Effects of variations in the Poissons ratio of the rock
mass (vr), are represented approximately by an equivalent
shear modulus of the rock mass (G*), defined as:
3v

G = Gr 1 + r

(118)

in which Gr = shear modulus of the elastic rock mass. For an


isotropic rock mass, the shear modulus is related to Er and vr by

FIGURE 45 Lateral loading of rock-socketed shaft (Carter and


Kulhawy 1992).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

63

Gr =

Er
2 (1 + vr )

(119)

Based on a parametric study using finite-element analysis,


it was found that closed-form expressions could be obtained
to provide reasonably accurate predictions of horizontal displacement (u) and rotation () at the head of the shaft for two
limiting cases. The two cases correspond to flexible shafts and
rigid shafts. The criterion for a flexible shaft is
D Ee

B G

2/7

(120)

For shafts satisfying Eq. 120, the response depends only


on the modulus ratio (Ee/G*) and Poissons ratio of the rock
mass (vr) and is effectively independent of D/B. The following closed-form expressions, suggested by Randolph (1981),
provide accurate approximations for the deformations of
flexible shafts:
H E
u = 0.50 e
G B G

H E
= 1.08 2 e
G B G

M E
+ 1.08 2 e
G B G

M E
+ 6.40 3 e
G B G

(121)
7

(122)

in which u = groundline deflection and = groundline rotation of the shaft.


Carter and Kulhawy (1992) reported that the accuracy of
the above equations is verified for the following ranges of parameters: 1 Ee/Er 106 and D/B 1.
The criterion for a rigid shaft is
D
E
0.05 e
G
B

(123)

and
Ee

G 100
2

2D

The accuracy of Eqs. 125 and 126 has been verified for the
following ranges of parameters: 1 D/B 10 and Ee/Er 1.
Shafts can be described as having intermediate stiffness
whenever the slenderness ratio is bounded approximately as
follows:
E
0.05 e
G

D E
< < e
B G

For the intermediate case, Carter and Kulhawy suggested


that the displacements be taken as 1.25 times the maximum
of either (1) the predicted displacement of a rigid shaft with
the same slenderness ratio (D/B) as the actual shaft or (2) the
predicted displacement of a flexible shaft with the same modulus ratio (Ee/G*) as the actual shaft. Values calculated in this
way should, in most cases, be slightly larger than those given
by the more rigorous finite-element analysis for a shaft of intermediate stiffness.
Carter and Kulhawy next considered a layer of soil of
thickness Ds overlying rock, as shown in Figure 46. The
analysis is approached by structural decomposition of the
shaft and its loading, as shown in Figure 46b. It was assumed
that the magnitude of applied lateral loading is sufficient to
cause yielding within the soil from the ground surface to the
top of the rock mass. The portion of the shaft within the soil
is then analyzed as a determinant beam subjected to known
loading. The displacement and rotation of point A relative to
point O can be determined by established techniques of structural analysis. The horizontal shear force (Ho) and bending
moment (Mo) acting in the shaft at the rock surface level can
be computed from statics, and the displacement and rotation
at this level can be computed by the methods described previously. The overall groundline displacements can then be
calculated by superposition of the appropriate parts.
Determination of the limiting soil reactions is recommended for the two limiting cases of cohesive soil in
undrained loading ( = 0) and frictional soil (c = 0) in drained
loading. Ultimate resistance for shafts in cohesive soils is
based on the method of Broms (1964a), in which the undrained

(124)

When Eqs. 123 and 124 are satisfied, the displacements of


the shaft will be independent of the modulus ratio (Ee/Er) and
will depend only on the slenderness ratio (D/B) and Poissons
ratio of the rock mass (vr). The following closed-form expressions give reasonably accurate displacements for rigid shafts:
H
2D
u = 0.4
G B B
H 2 D
= 0.3 2
G B B

M
2D
+ 0.3 2
G B B
M 2 D
+ 0.8 3
G B B

(125)

(126)

(127)

FIGURE 46 Rock-socketed shaft with overlying


layer (Carter and Kulhawy 1992).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

64

soil resistance ranges from zero at the ground surface to a


depth of 1.5B and has a constant value of 9su below this depth,
where su = soil undrained shear strength. For socketed shafts
extending through a cohesionless soil layer, the following limiting pressure suggested by Broms (1964b) is assumed:
pu = 3K p 'v
Kp =

1 + sin '
1 sin '

(128)
(129)

in which v' = vertical effective stress and ' = effective


stress friction angle of the soil. For both cases (undrained and
drained) solutions are given by Carter and Kulhawy (1992)
for the displacement, rotation, shear, and moment at point O
of Figure 46. The contribution to groundline displacement
and rotation from the loading transmitted to the rock mass
(Ho and Mo) is determined based on Eqs. 121 and 122 or
Eqs. 125 and 126 and added to the calculated displacement
and rotation at the top of the socket to determine overall
groundline response.
Application of the proposed theory is described by Carter
and Kulhawy (1992) through back-analysis of a single case involving field loading of a pair of rock-socketed shafts. The
method has not been evaluated against a sufficient database
of field performance, and further research is needed to assess
its reliability. The analysis was developed primarily for application to electrical transmission line foundations in rock,
although the concepts are not limited to foundations supporting a specific type of structure. The approach is attractive for
design purposes, because the closed-form equations can be
executed by hand or on a spreadsheet.
Carter and Kulhawy (1992) stated that the assumption of
yield everywhere in the soil layer may represent an oversimplification, but that the resulting predictions of groundline
displacements will overestimate the true displacements, giving a conservative approximation. However, the assumption
that the limit soil reaction is always fully mobilized may lead
to erroneous results by overestimating the load carried by the
soil and thus underestimating the load transmitted to the
socket. Furthermore, groundline displacements may be underestimated because actual soil resistance may be smaller
than the limiting values assumed in the analysis.

Zhang et al. (2000)


Zhang et al. (2000) extended the continuum approach to predict the nonlinear lateral load-displacement response of rocksocketed shafts. The method considers subsurface profiles
consisting of a soil layer overlying a rock layer. The deformation modulus of the soil is assumed to vary linearly with
depth, whereas the deformation modulus of the rock mass is
assumed to vary linearly with depth and then to stay constant
below the shaft tip. Effects of soil and/or rock mass yielding
on response of the shaft are considered by assuming that the

soil and/or rock mass behaves linearly elastically at small


strain levels and yields when the soil and/or rock mass reaction force p (force/length) exceeds the ultimate resistance pult
(force/length).
Analysis of the loaded shaft as an elastic continuum is accomplished using the method developed by Sun (1994). The
numerical solution is by a finite-difference scheme and incorporates the linear variation in soil modulus and linear
variation in rock mass modulus above the base of the shaft.
Solutions obtained for purely elastic responses are compared
with those of Poulos (1971) and finite-element solutions by
Verruijt and Kooijman (1989) and Randolph (1981). Reasonable agreement with those published solutions is offered
as verification of the theory, for elastic response.
The method is extended to nonlinear response by accounting for local yielding of the soil and rock mass. The soil and
rock mass are modeled as elastic, perfectly plastic materials,
and the analysis consists of the following steps:
1. For the applied lateral load H and moment M, the shaft
is analyzed by assuming the soil and rock mass are
elastic, and the lateral reaction force p of the soil and
rock mass along the shaft is determined by solution of
the governing differential equation and boundary conditions at the head of the shaft.
2. The computed lateral reaction force p is compared
with the ultimate resistance pult. If p > pult, the depth of
yield zy in the soil and/or rock mass is determined.
3. The portion of the shaft in the unyielded soil and/or
rock mass (zy z L) is considered to be a new shaft
and analyzed by ignoring the effect of the soil and/or
rock mass above the level z = zy. The lateral load and
moment at the new shaft head are given by:
Ho = H

zy
0

pult dz

M o = M + Hz y

zy
0

(130)
pult ( z y z ) dz

(131)

4. Steps 2 and 3 are repeated and the iteration is continued


until no further yielding of soil or rock mass occurs.
5. The final results are obtained by decomposition of
the shaft into two parts, which are analyzed separately, as illustrated previously in Figure 46. The section of the shaft in the zone of yielded soil and/or
rock mass is analyzed as a beam subjected to a distributed load of magnitude pult. The length of shaft in
the unyielded zone of soil and/or rock mass is analyzed as a shaft with the soil and/or rock mass behaving
elastically.
Ultimate resistance developed in the overlying soil layer
is evaluated for the two conditions of undrained loading
( = 0) and fully drained loading (c = 0). For fine-grained
soils (clay), undrained loading conditions are assumed and
the limit pressure is given by

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

65

pult = N p cu B
Np = 3 +

'
J
z+
z9
cu
2R

(132)
(133)

in which cu = undrained shear strength, B = shaft diameter, ' =


average effective unit weight of soil above depth z, J = a coefficient ranging from 0.25 to 0.5, and R = shaft radius. For
shafts in sand, a method attributed to Fleming et al. (1992)
is given as follows:
pult = K p2 'zB

(134)

where Kp = Rankine coefficient of passive earth pressure defined by Eq. 129. Ultimate resistance of the rock mass is
given by
pult = ( pL + max ) B

(135)

where max = maximum shearing resistance along the sides of


the shaft (e.g., Eq. 30 of chapter three) and pL = normal limit
resistance. The limit normal stress pL is evaluated using the
HoekBrown strength criterion with the strength parameters
determined on the basis of correlations to GSI. The resulting
expression was given previously as Eq. 113.
According to Zhang et al. (2000), a computer program was
written to execute this procedure. Predictions using the proposed method are compared with results of field load tests reported by Frantzen and Stratten (1987) for shafts socketed into
sandy shale and sandstone. Computed pile head deflections
show reasonable agreement with the load test results. The
method appears to have potential as a useful tool for foundations designers. Availability of the computer program is unknown. Programming the method using a finite-difference
scheme as described by Zhang et al. (2000) is also possible.
Discontinuum Models

A potential mode of failure for a laterally loaded shaft in rock


is by shear failure along joint surfaces. To et al. (2003) proposed a method to evaluate the ultimate lateral-load capacity
of shafts in rock masses with two or three sets of intersecting
joints. The analysis consists of two parts. In the first part, the
block theory of Goodman and Shi (1985) is used to determine if possible combinations of removable blocks exist that
would represent a kinematically feasible mode of failure.
In the second part, the stability of potentially removable
combinations of blocks or wedges is analyzed by limit equilibrium. Both steps in the analysis require careful evaluation
of the joint sets, in terms of their geometry and strength properties. Although the method is based on some idealized
assumptions, such as equal joint spacing, and it has not been
evaluated against field or laboratory load tests, it provides a
theoretically based discontinuum analysis of stability in
cases where this mode of failure requires evaluation.

Discussion of Analytical Models for Laterally


Loaded Sockets

Each of the analytical methods described above has advantages and disadvantages for use in the design of rock-socketed
shafts for highway bridge structures. The greatest need for
further development of all available methods is a more thorough database of load test results against which existing
theory can be evaluated, modified, and calibrated.
The simple closed-form expressions given by Carter and
Kulhawy (1992) represent a convenient, first-order approximation of displacements and rotations of rock-socketed
shafts. Advantages include the following:
Predicts lateral displacements under working load
conditions,
Requires a single material parameter (rock mass
modulus),
Provides reasonable agreement with theoretically
rigorous finite-element analysis, and
Is the easiest method to apply by practicing design
engineers.
Limitations include:
Does not predict the complete lateral load-displacement
curve,
Elastic solution does not provide shear and moment distribution for structural design,
Does not account for more than one rock mass layer,
Does not account directly for nonlinear MEI behavior
of reinforced-concrete shaft, and
Does not account for interaction between axial and lateral
loading and its effects on structural behavior of the shaft.
The method can be best used for preliminary design; for
example, establishing the initial trial depth and diameter of
rock-socketed shafts under lateral and moment loading. For
some situations, no further analysis may be necessary. Final
design should be verified by field load testing.
The method of Zhang et al. (2000) provides a more rigorous
continuum-based analysis than that of Carter and Kulhawy.
The tradeoff is that more material parameters are required as
input. Variation of rock mass modulus with depth is required.
To fully utilize the nonlinear capabilities, the HoekBrown
yield criterion parameters are required, and these are based on
establishing the RMR and/or GSI. The method is best applied
when a more refined analysis is required and the agency is willing to invest in proper determination of the required material
properties. Advantages include:
Predicts the full, nonlinear, lateral load-deformation response;
Accounts for partial yield in either the rock mass or the
overlying soil (more realistic);

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

66

Is based on well-established rock mass and soil properties;


Is verified against rigorous theory, for elastic range; and
Provides shear and moment distribution for structural
design.
Limitations include:
Requires numerical (computer) solution, not currently
available commercially;
Requires a larger number of rock mass material parameters;
Currently is limited to two layers (one soil and one rock
mass layer, or two rock mass layers); and
Nonlinear MEI behavior of reinforced-concrete shaft
is not accounted for explicitly; requires iterative analyses with modified values of EI.
The most rigorous analytical methods based on a continuum
approach are FEM. When implemented by competent users,
FEM analysis can account for the shaft, soil, and rock mass
behaviors more rigorously than the approximate methods described herein, but FEM analyses are not suitable for routine
design of foundations in most cases. First, the results are only
as reliable as the input parameters. In most cases the material
properties of the rock mass are not known with sufficient reliability to warrant the more sophisticated analysis. Second, the
design engineer should have the appropriate level of knowledge
of the mathematical techniques incorporated into the FEM
analyses. Finally, the time, effort, and expense required for
conducting FEM analyses are often not warranted. For very
large or critical bridge structures, sophisticated FEMs may be
warranted and the agency might benefit from the investment
required in computer codes, personnel training, and field and
laboratory testing needed to take advantage of such techniques.
Subgrade reaction methods, as implemented through the
p-y curve method of analysis, offer some practical advantages
for design. These include:

Considering that the p-y method is currently being used


extensively by most state DOTs, effort should be made to
address its present limitation by research aimed at better establishing methods to specify appropriate p-y curves in rock.
Full-scale field load testing with instrumentation is the only
known method to verify p-y curves. Research conducted for
this purpose would provide an opportunity to evaluate and
calibrate other proposed analytical methods; for example,
those of Carter and Kulhawy (1992) and Zhang et al. (2000)
and for development of new models. Recommendations for
research are discussed further in chapter five. The research
programs sponsored by the North Carolina and Ohio DOTs
illustrate the type of approach that is useful for advancing all
of the available methods of analysis. In addition to providing
improved criteria for p-y curve modeling, the load test results
reported by Gabr et al. (2002) and Liang and Yang (2006)
can be used to evaluate each others models and the SW and
continuum models described in this chapter.
In summary, a range of analytical tools are available to
foundation designers to consider rock sockets under lateral
and moment loading. These include simple, closed-form
equations requiring a small number of material properties
(Carter and Kulhawy 1992). A more rigorous model that
predicts the complete nonlinear response but requires more
material properties is also available (Zhang et al. 2000).
Highly sophisticated numerical models requiring extensive
material properties and appropriate expertise (FEM analysis)
exist and may be appropriate for larger projects. The p-y
method of analysis is attractive to designers, as evidenced by
its wide use; however, considerable judgment is required in
selection of p-y curve parameters. All of the currently available methods suffer from a lack of field data for verification
and are best applied in conjunction with local and agency experience, thorough knowledge of the geologic environment,
and field load testing.

STRUCTURAL ISSUES

Predicts the full, nonlinear lateral load-deformation


response;
Can incorporate multiple layers of soil and/or rock;
Accounts for nonlinear MEI behavior of reinforcedconcrete shaft;
Provides structural analysis (shear, moment, rotation,
and displacement) of the drilled shaft;
Accounts for the effects of axial compression load on
the structural behavior of the shaft; and
Can be implemented easily on a desktop computer with
available software.
The principal limitations are:
Lack of a strong theoretical basis for p-y curves and
Requires back analysis of instrumented load tests to
verify and validate p-y curves; such verification is currently lacking or limited to a few cases.

Twenty of the questionnaire responses indicated that structural design of drilled shaft foundations is carried out by engineers in the Bridge Design or Structures Division of their
state DOTs. Three states indicated that structural design is a
joint effort between the Geotechnical and Structural/Bridge
Divisions. One DOT indicated that structural design is done
by the Geotechnical Branch. All of the states responding to
the structural design portion of the questionnaire stated that
the AASHTO LRFD Bridge Design Specifications are followed for structural design of drilled shafts. Three states also
cited the ACI Building Code Requirements for Structural
Concrete.
Barker et al. (1991) discussed the structural design of
reinforced-concrete shafts and have several design examples
illustrating the basic concepts. ONeill and Reese (1999) also
covered the general aspects of reinforced-concrete design for

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

67

drilled shafts, for axial compression loading and flexure,


citing as primary references the 1994 AASHTO LRFD Bridge
Design Specifications (1st edition) and the 1995 ACI Building
Code Requirements for Structural Concrete (ACI 318-94).
Both the AASHTO and ACI codes have since been revised
(AASHTO in 2004 and ACI in 2002); however, there are no
major differences that would change the structural design of
drilled shafts. According to the survey, all of the states are
designing in accordance with the AASHTO LRFD Bridge
Design Specifications. At the time of this writing, Section 10
(Foundations) of the draft 2006 Interim AASHTO LRFD
Bridge Design Specifications was available for reference.
However, the other sections of the 2006 Interim specifications were not available and so comments pertaining to Section 5 are referenced to the 2004 edition. Only issues of structural design pertaining specifically to rock-socketed drilled
shafts are addressed here.

P
M

M
V
Top of
Column

V
y

Equivalent
Fixed-end
Column

Deep
Foundation

depth of fixity
FIGURE 47 Depth of fixity for equivalent fixed-end column.

General Issues

Section 10.8.3.9 (Shaft Structural Resistance) of the 2006


Interim AASHTO LRFD Bridge Design Specifications states
that
The structural design of drilled shafts shall be in accordance with
the provisions of Section 5 for the design of reinforced concrete.

This language makes it clear that drilled shaft structural


design is subject to the same provisions as other reinforcedconcrete members. The designer must then determine
whether the shaft is a compression member or a member subjected to compression and flexure (beam column). Article
10.8.3.9.3 states the following:
Where the potential for lateral loading is insignificant, drilled
shafts may be reinforced for axial loads only. Those portions of
drilled shafts that are not supported laterally shall be designed as
reinforced-concrete columns in accordance with Article 5.7.4.
Reinforcing steel shall extend a minimum of 10 ft below the
plane where the soil provides fixity.

The commentary accompanying Article 10.8.3.9.3 states further that:


A shaft may be considered laterally supported: below the zone
of liquefaction or seismic loads, in rock, or 5.0 ft below the
ground surface or the lowest anticipated scour elevation. . . . .
Laterally supported does not mean fixed. Fixity would occur
somewhere below this location and depends on the stiffness of
the supporting soil.

The language in this provision could be improved by providing a definition of fixity. Fixity is defined by Davisson
(1970) for piles under lateral loading as the depth below
groundline corresponding to the fixed base of an equivalent
free-standing column; that is, a column for which the top
deflection and rotation would be the same as that of a column
supported by the embedded deep foundation (Figure 47).
Approximate equations are given by Davisson for establishing

depth of fixity based on the depth of the foundation and a


relative stiffness factor that depends on the flexural rigidity
of the pile and the subgrade modulus of the soil or rock. Interviews with state DOT engineers indicated that different
criteria for establishing depth of fixity are being applied. One
state DOT defines fixity as the depth at which LPILE analysis shows the maximum moment, whereas another defines
fixity as the depth at which LPILE shows zero lateral deflection. In Section 12 of Bridge Design Aids (1990), the Massachusetts Highway Department) describes a rigorous approach involving use of the program LPILE (or other p-y
analysis) to establish a depth of fixity as defined in Figure 47.
For the given soil/rock profile, approximate service loads
are applied to the Top of Column (Figure 47). Shear and
moment are applied as separate load cases and the resulting
lateral deflections and rotations at the top of the column are
designated as follows:
V = deflection due to shear (V)
M = deflection due to moment (M)
V = rotation due to shear (V)
M = rotation due to moment (M).
Equivalent column lengths are then calculated using the
following analytical expressions for each loading case. The
four resulting values of L should be approximately equal and
the average value can be taken as a reasonable approximation of the equivalent fixed-end column length. Depth of fixity corresponds to the portion of the fixed-end column below
groundline.
LV

3V ( EI )
=

2 M ( EI )
LM =

(136)
2

Copyright National Academy of Sciences. All rights reserved.

(137)

Rock-Socketed Shafts for Highway Structure Foundations

68

LV

2V ( EI )
=

LM =

M ( EI )
M

Moment Transfer
2

(138)

(139)

The principal use of depth of fixity is to establish the


elevation of equivalent fixed-end columns supporting the
superstructure, thus enabling structural designers to uncouple the foundations from the superstructure for the purpose of
structural analysis and design of the bridge or other structure.
Structural modeling of the superstructure with equivalent
fixed-end columns is also used to establish the column loads.
These column loads are then used to analyze the drilled shaft
foundations by applying them to the top of the actual column, which is continuous with the foundation (left side of
Figure 47) using p-y analysis. As described at the beginning
of this chapter, these analyses may be done by either the GD
or Bridge offices, but the soil and rock parameters are provided by GD. The p-y analysis gives the maximum moment
and shear that are used in the reinforced-concrete design. Use
of software such as LPILE, COM624, or other programs
is thus seen to be an integral tool in both the geotechnical
and structural design of drilled shafts for bridges or other
transportation structures. As noted previously, AASHTO
specifications define the strength limit state for lateral loading only in terms of foundation structural resistance. Lateral
deflections as predicted by p-y analyses are used as a design
tool to satisfy service limit state criteria.
The concept of fixity also has implications for reinforcing
steel requirements of drilled shafts. According to Article
10.8.3.9.3, as cited earlier, if a drilled shaft designed for axial
compression extends through soil for a distance of at least
3 m (10 ft) beyond fixity before entering into rock, the rocksocketed portion of the shaft does not require reinforcement.
This provision would also limit the need for compression
steel in rock sockets to a maximum depth of 3 m below fixity. Exceptions to this are shafts in Seismic Zones 3 and 4,
for which Article 5.13.4.6.3d states that for cast-in-place
piles, longitudinal steel shall be provided for the full length
of the pile.
Some state DOTs use permanent steel casing in the top portion of drilled shafts or, in many cases, down to the top of rock.
Permanent casing is not mentioned in the 2006 Interim specifications, but the 2004 specifications included the following
statement: Where permanent steel casing is used and the shell
is smooth pipe greater than 0.12 in. thick, it may be considered
to be load-carrying. Allowance should be made for corrosion.
A few states indicated that questions arise in connection
with relatively short sockets in very hard rock. The questions
pertain to moment transfer, development length of steel reinforcing, and apparently high shear loads resulting from
high moment loading.

Rock sockets subjected to high lateral and/or moment loading


require a minimum depth of embedment to transfer moment
to the rock mass and to satisfy minimum development length
requirements for reinforcing steel. The mechanism of moment
transfer from a column to the rock is through the lateral
resistance developed between the concrete shaft and the rock.
The resistance depends on many of the factors identified
previously, primarily strength and stiffness of the rock mass
and flexural rigidity of the shaft. When the strength and modulus of the rock mass are greater than that of the concrete
shaft, the question may arise, why excavate such high quality rock and replace it with lower strength concrete? The only
means to transfer moment into the rock mass is through a
properly designed shaft with the dimensions, strength, and
stiffness to transmit the design moment by the assumed
mechanisms of lateral resistance. In some situations where
high strength rock mass is close to the ground surface, shaft
size may be governed by structural considerations rather than
by geotechnical capacity.
For some relatively short, stubby shafts in hard rock,
socket length could be governed by the required development length of longitudinal reinforcing bars. Article 5.11 of
the AASHTO LRFD Bridge Design Specifications (2004)
specifies basic tension development lengths for various bar
sizes as a function of steel and concrete strengths. Because
the bars will be stressed to their maximum values at the
points where maximum moments occur, the distance between
the point of maximum moment and the bottom of the socket
must be at least equal to the required development length.
As an example, for No. 18 bars, assuming fy = 414 MPa
(60 ksi) and fc' = 27.6 MPa (4 ksi), basic development length
is 267 cm (105 in. or 8.75 ft). Although this is not often the
governing factor for socket length, it should be checked.

Shear

Some designers commented on cases where p-y analysis of


laterally loaded rock-socketed shafts resulted in unexpectedly
high values of shear and whether the results were realistic.
In particular, when a rock socket in relatively strong rock is
subjected to a lateral load and moment at its head, values of
shear near the top of the socket may be much higher than the
applied lateral load. This result would be expected given the
mechanism of moment load transfer. When the lateral load
has a high moment arm, such as occurs in an elevated structure, the lateral load transmitted to the top of the drilled shaft
may be small or modest, but the moment may be relatively
large. The principal mechanism of moment transfer from the
shaft to the rock mass is through the mobilized lateral resistance. If a large moment is transferred over a relatively short
depth, the lateral resistance is also concentrated over a relatively short length of the shaft and results in shear loading
that may be higher in magnitude than that of the lateral load.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

69

There is some question, however, whether high values of


shear predicted by p-y methods of analysis for such cases
exist in reality or are artifacts of the analysis. One designer suggested that the structural model of the shaft does not account
properly for shear deformation, resulting in unrealistically high
shear values. The topic requires further investigation.

where Avs = area of shear reinforcement, s = longitudinal (vertical) spacing of the ties or pitch of the spiral, and d = effective
shear depth. For a circular cross section this can be taken as

In some cases, the magnitude of shear must be addressed


in the reinforced-concrete design, primarily in the use of
transverse reinforcement. According to AASHTO LRFD
Bridge Design Specifications (2004), the minimum amount
of spiral reinforcement to satisfy the requirements for compression is governed by

The need for additional transverse reinforcement, beyond


that required for compression, can be determined by Eq. 141.
For the majority of rock-socketed shafts, the transverse reinforcement required to satisfy compression criteria (Eq. 140)
combined with the shear resistance provided by the concrete
(Eq. 142) will be adequate to resist the factored shear loading without the need for additional transverse reinforcement.
However, in cases where high lateral load or moment are to
be distributed to the ground over a relatively small distance;
for example, a short stubby socket in high-strength rock,
factored shear forces may be high and the shaft dimensions
and reinforcement may be governed by shear. In these cases,
the designer is challenged to provide a design that provides
adequate shear resistance without increasing the costs excessively or adversely affecting constructability by constricting
the flow of concrete.

Ag f '
s = 0.45 1 c
Ac f y

(140)

in which s = ratio of spiral reinforcement to total volume of


concrete core, measured out-to-out of spirals; Ag = gross
(nominal) cross-sectional area of concrete; and Ac = crosssectional area of concrete inside the spiral steel. When shear
occurs in addition to axial compression, the section is then
checked by comparing the factored shear loading with the
factored shear resistance, given by
Vr = Vn = (Vc + Vs )

(141)

in which Vr = factored shear resistance, Vn = nominal shear


resistance, Vc = nominal shear resistance provided by the
concrete, Vs = nominal shear resistance provided by the transverse steel, and = resistance factor = 0.90 for shear. The
nominal shear strength provided by the concrete is given (in
U.S. customary units) by:

Pu
Vc = 2 1 +
2, 000 Ag

f c' Av

(142)

or, when axial load is zero,


Vc = 2 f c' Av

(143)

where Pu = factored axial load and Av = area of concrete in the


cross section that is effective in resisting shear. For a circular
section this can be taken as
B B
Av = 0.9 B + ls

2

(144)

in which B = shaft diameter and Bls = diameter of a circle


passing through the center of the longitudinal reinforcement.
The nominal shear strength provided by transverse reinforcement is given by
Vs =

Avs f y d
s

(145)

B B
d 0.9 + ls

2

(146)

To handle high shear loading in the reinforced-concrete


shaft, the designer has several options: (1) increase the shaft
diameter, thus increasing the area of shear-resisting concrete;
(2) increase the shear strength of the concrete; or (3) increase
the amount of transverse reinforcing, either spiral or ties, to
carry the additional shear. Each option has advantages and
disadvantages.
Two variables that can be adjusted to increase shear resistance are concrete 28-day compressive strength, f c' , and
shaft diameter, B. Increasing the concrete strength can be a
cost-effective means of increasing shear strength. For example, increasing fc' from 27.6 MPa to 34.5 MPa (4000 psi to
5000 psi) yields a 12% increase in shearing resistance. Increasing the diameter of a rock socket can add considerably
to the cost, depending on rock type, drillability, socket depth,
etc. Rock of higher strength, which is likely to coincide with
the case when shear is critical, can be some of the most expensive rock to drill. However, increasing the diameter can
provide other benefits that may offset additional costs, such
as reducing the congestion of reinforcement steel (improved
constructability), increasing axial and bending capacity, and
further limiting displacements.
Shear strength of the shaft can also be increased by
providing additional transverse reinforcement in the form of
either spiral or ties. From Eq. 145, this can be achieved by increasing the size of transverse reinforcement or by decreasing the pitch(s). Constructability can be affected when bar
spacings are too small to allow adequate flow of concrete.
One aspect of reinforced-concrete behavior in shear that
is not taken into account in any building code is confining

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

70

stress. Shear capacity of concrete is increased at higher


confining stress and deep foundations are subjected to significant confinement, especially when they are embedded
in rock. This is a topic that warrants research but has yet to
be investigated in a meaningful way that can be applied to
foundation design.
Axial

When lateral loading is not significant, structural design of


concrete shafts must account for axial compression or tension (e.g., uplift) capacity. For shafts designed for significant load transfer at the base, compression capacity of the
reinforced-concrete shaft could be less than that of the rock
bearing capacity. In high-strength intact rock, compressive
strength of the shaft may be the limiting factor. For design of
reinforced-concrete columns for axial compression, the
AASHTO-factored axial resistance is given by

field. However, the ability of analytical methods to account


properly for rock mass response and rockstructure interaction has not developed to the same level as methods used for
deep foundations in soil.
The survey shows that most state DOTs use the program
COM624 or its commercial version LPILE for design of
rock-socketed shafts. Review of the p-y curve criteria currently built into these programs for modeling rock mass response shows that they should be considered as interim and
that research is needed to develop improved criteria. Some of
this work is underway and research by North Carolina (Gabr
et al. 2002), Ohio (Liang and Yang 2006), Florida (McVay
and Niraula 2004), and Ashour et al. (2001) is described. All
of these criteria are in various stages of development and are
not being applied extensively.

SUMMARY

Models based on elastic continuum theory and developed


specifically for rock-socketed shafts have been published.
Two methods reviewed in this chapter are the models of
Carter and Kulhawy (1992) and Zhang et al. (2000). Advantages and disadvantages of each are discussed and compared
with p-y methods of analysis. These models are most useful
as first-order approximations of shaft lateral displacements
for cases where the subsurface profile can be approximated
as consisting of one or two homogeneous layers. For example, when a preliminary analysis is needed to develop trial
designs that will satisfy service limit state deflection criteria,
the method of Carter and Kulhawy can provide convenient
solutions that can be executed by means of spreadsheet
analysis. A disadvantage of these methods is that they do not
directly provide solutions to maximum shear and moment,
parameters needed for structural design, and they do not incorporate directly the nonlinear properties of the reinforcedconcrete shaft.

Lateral loading is a major design consideration for transportation structures and in many cases governs the design of
rock-socketed drilled shafts. Design for lateral loading must
satisfy performance criteria with respect to (1) structural resistance of the reinforced-concrete shaft for the strength limit
state and (2) deflection criteria for the service limit state.
Analytical methods that provide structural analysis of deep
foundations while accounting for soilstructure interaction
have, therefore, found wide application in the transportation

Structural issues associated with rock-socketed shafts are


reviewed. The concept of depth of fixity is shown to be a useful analytical tool providing a link between geotechnical and
structural analysis of drilled shafts. A method for establishing depth of fixity is presented and its use in the design
process is described. Other issues identified by the survey,
including high shear in short sockets subjected to high
moment loading and its implications for reinforced-concrete
design, are addressed.

Pr = 0.85 [ 0.85 f c' ( Ag Ast ) + f y Ast ]

(147)

in which Pr = factored axial resistance, with or without flexure; = resistance factor (0.75 for columns with spiral transverse reinforcement, 0.70 for tied transverse reinforcement);
fc' = strength of concrete at 28 days; Ag = gross area of the
section; Ast = total area of longitudinal reinforcement; and
fy = specified yield strength of reinforcement. One source of
uncertainty is that the design equations given here are for
unconfined reinforced-concrete columns. The effect of confinement provided by rock on the concrete strength is not
easy to quantify, but increases the strength compared with
zero confinement, and warrants further investigation.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

71

CHAPTER FIVE

CONSTRUCTION AND FIELD TESTING

SCOPE

Construction, inspection, post-construction integrity testing,


and load testing of drilled shafts are related directly to design and performance. These activities are carried out in the
field and depend on the skill and experience of contractors,
technicians, inspection personnel, and engineers. In this
chapter an overview is presented of construction methods for
rock sockets. Methods for load testing of rock-socketed
shafts are reviewed, including several innovative methods
that have made load testing more accessible to state transportation agencies. Illustrative examples demonstrate how
load testing can contribute to the economical design of rock
sockets. Constructability issues identified by the survey questionnaire are discussed, and practices that can lead to quality
construction are identified. Current practice for inspection
and quality assurance methods for rock-socketed shafts are
also reviewed and discussed. Finally, special geologic conditions that pose unique challenges for design and construction
of rock sockets are described, and approaches for using rock
sockets successfully in such environments are identified.

CONSTRUCTION OF ROCK SOCKETS

The art and science of drilled shaft construction are as important to the success of a bridge foundation project as are
the analytical methods used to design the shafts. Construction of shafts in rock can be some of the most challenging and
may require special expertise and equipment. Experience
demonstrates that the key components of success are: (1) adequate knowledge of the subsurface conditions, for both
design and construction; (2) a competent contractor with the
proper equipment to do the job; and (3) a design that takes
into account the constructability of rock sockets for the particular job conditions. Publications that cover drilled shaft
construction methods include Greer and Gardner (1986) and
ONeill and Reese (1999). Aspects of construction that are
related to rock sockets are reviewed herein.

Drilling Methods and Equipment

Most rock-socketed shafts are excavated using rotary drilling


equipment. A rotary drill may be mechanically driven or use
hydraulic motors. Mechanically driven rigs deliver power to a
stationary rotary table that rotates a kelly bar to which excavation tools are attached. Mechanically driven rigs can be

truck-mounted or attached to a crane (Figure 48). Hydraulic


drilling rigs are equipped with hydraulic motors that can be
moved up and down the mast and are usually truck or crawler
mounted. Smaller hydraulic units can be mounted on an excavator. Hydraulic drilling rigs with significantly increased
power have appeared in the North American market in recent
years. Drilling in rock, especially hard rock, generally requires
machines with more power than for drilling in soil. Equipment
with higher torque ratings and additional power has given
more contractors the capability to install rock-socketed shafts
than existed previously. This is a positive development for the
U.S. market in that it promotes competition and expands the
base of experienced contractors for rock-socket construction.
Equipment developed in Europe and now being used by
some North American contractors uses hydraulic rams configured to rotate or oscillate (rotate back and forth) a steel
casing into the ground (Figure 49). Soil or rock is excavated
from inside the casing using a hammergrab, a percussion tool
that breaks and removes soil or rock. In most cases the rotator or oscillator is bolted to a crane for stability under the
large torque that must be developed. The crane can also provide hydraulic power to operate the rams that turn the casing.
Rotators have the capability to cut through high-strength
rock in the range of 100150 MPa (1522 ksi) depending on
the degree of fracturing (J. Roe, Malcolm Drilling, personal
communication, Oct. 3, 2005). A 3-m-diameter oscillator such
as the one shown in Figure 49 is generally limited to cutting
through weaker rock with strength less than 100 MPa. The
lead casing on the oscillator must have teeth set in opposite
directions to cut back and forth. Both methods are efficient
in penetrating large cobbles and boulders, a situation common to glacial till deposits and cemented sands and gravels.

Rock Cutting Tools

Selecting the proper cutting tool depends on many variables,


including rock mass properties (strength, hardness, and structure), type of drilling machine, socket depth and diameter, condition and cost of the tools, operator skill, previous experience
in similar conditions, and judgment. There are no absolute
rules and different contractors may take a completely different
approach when faced with similar conditions. New tools and
innovations are constantly being introduced. Following is a
summary of some of the most common cutting tools used for
rock-socket construction.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

72

FIGURE 50 Rock auger with drag bit and bullet-shaped cutting


teeth.

FIGURE 48 Intact rock core removed using crane-mounted


rotary drill and core barrel.

When relatively stiff soil or weak rock cannot be penetrated efficiently with typical soil drilling tools (e.g., open
helix augers), most contractors will attempt to use a rock
auger. Rock augers are manufactured from thicker metal
plate than soil augers and have cutting teeth. The teeth may
be of the drag bit type, which are effective in cutting rock but
wear rapidly and must be replaced frequently. As a rule of
thumb, these types of teeth are limited to cutting rock of compressive strength up to approximately 48 MPa (7,000 psi), at
which point they dull quickly. Conical-shaped teeth made of
tungsten carbide or other alloys depend on crushing the rock
and are more durable than drag bits, but require considerable
downward force (crowd) to be effective. Figure 50 shows a
rock auger with both types of teeth, to exploit both mechanisms

FIGURE 49 Casing oscillator and hammergrab tool.

of cutting and crushing. Rock augers may be stepped or


tapered so that the initial penetration into rock requires less
torque and crowd, or the socket may be drilled first by
a smaller diameter tool such as the one shown in Figure 51,
followed by a larger diameter auger. This releases some of
the confinement and causes less wear and tear on the drilling
tools. Replacement or reconditioning of rock auger teeth can
be a major contractor cost, especially in highly abrasive rock.
Self-rotating cutter bits combine a highly efficient cutting
mechanism with the durability of some conical bits. A rock
auger with self-rotating cutters, for excavating the face of the
socket, and conical bits directed outward is shown in Figure
52. A contractor using this auger reported penetration rates
two to three times higher than with conventional rock augers
and in very hard (100 MPa or 15,000 psi) rock.
At some combination of rock strength and socket diameter rock augers are no longer cost-effective. One contractor

FIGURE 51 Small diameter rock auger for creating a pilot hole.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

73

FIGURE 52 Rock auger with conical teeth and rotating cutters


(Courtesy : V. Jue, Champion Equipment, Inc.).

interviewed for this study stated that, for socket diameters up


to approximately 1.8 m and rock strength up to 70 MPa
(10,000 psi), initial cost estimates are based on the assumption that rock augers will be used. If the combination of
socket diameter and rock strength exceeds those values, the
job is bid on the assumption that rock will be cored. Of
course, these rule-of-thumb criteria are subject to change on
the basis of rock mass characteristics, experience, etc., and
will vary between contractors. Use of a single parameter,
such as uniaxial compressive strength of rock, does not capture all of the variables that determine penetration rates for a
given set of conditions.
Coring is a widely used method when rock augers are no
longer feasible. The basic concept is that coring reduces the
volume of rock that is actually cut by the teeth. A simple configuration consists of a single cylindrical barrel with cutting
teeth at the bottom edge (Figure 53). The teeth cut a clearance
on the inside and outside of the barrel that is sufficient for removing cuttings and extraction of the core barrel. The core
may break off at a discontinuity or it may require use of a rock

FIGURE 53 Typical single wall core barrel.

FIGURE 54 Welding roller bits on a 4-m-diameter doublewalled core barrel.

chisel, a metal tool that is wedged between the barrel and the
rock to fracture the core. The core will usually jam into the
barrel and can be lifted out of the hole and then removed by
hammering the suspended barrel (see Figure 48). If the rock is
highly fractured, the core barrel may be removed, followed by
excavation of the fractured rock from the hole. For deep sockets
or for harder rock, double wall core barrels may be used. The
outer barrel is set with teeth, typically roller bits (Figure 54),
while the core is forced into the inner barrel. Compressed air
is circulated between the barrels to remove cuttings.
For very high strength rock (qu 100 MPa) there are few
tools that will excavate efficiently. In these rocks, however,
even a small penetration can provide high axial, and in some
cases lateral, resistance. A shot barrel, in which hard steel
shot is fed into the annular space between the double walls
of the core barrel, may work in such conditions. Grinding action of the shot excavates the rock and water is circulated for
cooling the shot.
Excavation rates with core barrels are typically slow. Although coring may be cost-effective because of the foundation performance benefits achieved, careful attention should
be given to avoiding overly conservative designs that significantly increase the cost of drilled shafts made by unnecessary coring into rock.
Hard rock can also be excavated using downhole hammer
bits. The tool shown in Figure 55 has an array of button-bit
hammers (called a cluster drill) operated independently by
compressed air. Air pressure also lifts the cuttings which are
collected in a calyx basket. On the tool shown in Figure 55,
some of the bits can be rotated outward to create a larger
diameter socket (under reaming) than the casing, and then
retracted to remove the bit. This allows a casing to be installed
directly behind the bit during drilling. Downhole hammers and
cluster drills are generally expensive and require large air compressors to operate. Most contractors will rent this equipment
when needed, which is only cost-effective in very hard rock.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

74

FIGURE 55 Downhole hammer tool for drilling in hard rock.

A technique used for drilling large rock sockets at the


RichmondSan Rafael Bridge (Byles 2004) is reverse circulation drilling with a pile top rig. The unit consists of two
main components. A top unit (Figure 56) is fixed to the top
of a steel casing. The bottom hole assembly (Figure 57) is
a drill bit lowered to the bottom of the hole through a casing,
submerged in water or other drilling fluid. The bit is operated

hydraulically through lines extending from the top unit


(sometimes called a rodless drill), which is fixed to the top of
the casing. Alternatively, a drill rod may be used to transmit
torque from the top unit to the bit. The bit has a central hollow orifice connected to a flexible line extending back up to
the top unit. During drilling, a vacuum pump or air lift is used
to draw the drilling fluid with the cuttings upward to a cleaning plant, from where it is circulated back into the hole. The
unit shown in Figure 56 was used to drill 3.35-m-diameter
rock sockets in Franciscan Formation sandstone and serpentinite. Some manufacturers are now producing reverse
circulation units that can be installed on a conventional rotary
hydraulic drilling rig to provide similar capability, at a
smaller diameter. It is likely that these units will become
more common in North America for rock-socket drilling
(D. Poland, Anderson Drilling, personal communication,
Aug. 2, 2005). Reverse circulation drilling can also be carried out with any type of rotary drill rig equipped with a hollow Kelly bar (drill stem) that allows circulation of the
drilling fluid from the cutting surface up through the bar.

FIELD LOAD TESTING

The most direct method to determine the performance of fullscale rock-socketed drilled shafts is through field load testing. Clearly there have been advances in engineers ability to
predict rock-socket behavior. However, there will always be
sources of uncertainty in the applicability of analysis methods, in the rock mass properties used in the analysis, and with
respect to the unknown effects of construction. Load testing
provides direct measurement of load displacement response
for the particular conditions of the test foundation, and can
also provide data against which analytical models can be
evaluated and calibrated.

Objectives

Field load testing may be conducted with different objectives


and this should determine the scope of testing, type of tests,
and instrumentation. A partial listing of valid reasons for
transportation agencies to undertake load testing of rocksocketed shafts includes:

FIGURE 56 Toredo T40-4 pile top unit being placed over


casing for reverse circulation drilling, RichmondSan Rafael
Bridge (California) (Byles 2004).

Confirm design assumptions,


Evaluate rock resistance properties,
Evaluate construction methods,
Reduce foundation costs, and
Research aimed at evaluating or improving design
methods.

More than one of these objectives can sometimes be


achieved. For example, load tests conducted primarily for
confirmation of design assumptions (proof test) for a particular project can be useful to researchers by contributing additional data for evaluating empirical correlations proposed

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

75

FIGURE 57 Shrouded bottom hole assembly lifted for placement through


the top unit (Byles 2004).

for design. Load tests carried to ultimate capacity of the shaft


are especially valuable not only to the agency conducting the
test or for the specific bridge project, but to the entire deep
foundation engineering community.
The costs of conducting field load tests should be offset
by its benefits. The most obvious costs include the dollar
amount of contracts for conducting testing. Other costs that
are not always as obvious include construction delays, delays
in design schedule, and DOT person hours involved in the
testing. Direct cost benefits may be possible if the testing
leads to more economical designs. This requires testing prior
to or during the design phase. Numerous case histories in the
literature show that load testing almost always leads to
savings. Lower factors of safety and higher resistance factors
are allowed by AASHTO for deep foundation design when
a load test has been conducted.
Other benefits may not be so obvious or may occur over
time. Construction of the test shaft provides the DOT and
all subsequent bidders with valuable information on constructability that can result in more competitive bids. Refinement in design methods resulting from information gained by
load testing has economic benefits on future projects.
Load test results provide the most benefit when they are
accompanied by high-quality subsurface characterization.
Knowledge of site stratigraphy, soil and rock mass properties,
site variability, and groundwater conditions are essential for
correct interpretation of load test results. The ability to apply
load test results to other locations is enhanced when subsurface conditions can be compared on the basis of reliable data.
Construction factors and their potential effects on shaft
behavior should be considered when using load test results

as the basis for design of productions shafts. Items such as


construction method (casing, slurry, dry), type of drilling
fluid, cleanout techniques, and others may have influenced
the behavior of the test shaft. If possible, the construction
methods anticipated for production shafts should be used to
construct test shafts.

Axial Load Testing

Conventional Axial Load Testing


Until the early 1990s the most common procedure for conducting a static axial compression load test on a deep foundation followed the ASTM Standard Method D1143, referred to
herein as a conventional axial load test. Several load application methods are possible, but the most common involves
using either (1) a hydraulic jack acting against a reaction
beam that is anchored against uplift by piles or (2) a loading
platform over the pile top on which dead load is placed. Six
states indicated that they have conducted conventional axial
load tests on rock-socketed shafts. Conduct and interpretation
of axial compression and uplift load tests specifically for drilled
shafts is discussed in detail by Hirany and Kulhawy (1988).
Axial load tests may be conducted for the purpose of confirming the design load for a specific project, in which case
it is typical to load the shaft to twice the anticipated design
load to prove the shaft can support the load with an acceptable settlement (a proof load test). This type of test is normally conducted under the construction contract and does not
yield a measured ultimate capacity, unless the shaft fails, in
which case the design must be adjusted. Proof tested shafts
normally are not instrumented except to measure load and
displacement at the head of the shaft. When the objective of

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

76

testing is to gain information on behavior of the shaft in terms


of load transfer, the shaft should be instrumented to determine the distribution of axial load as a function of depth and
as a function of axial deformation.
Common types of instrumentation for measuring axial
load and deformation at specific points along the length of
the shaft include sister bars and telltales. A sister bar is a section of reinforcing steel, typically 1.2 m in length, with a
strain gage attached in the center. Either vibrating wire or
electrical resistance-type gages can be used. The sister bar is
tied to steel of the reinforcing cage and its lead wires are
routed to the surface, where they are monitored by a computer-controlled data acquisition unit. The gage signals are
converted to strain, which is assumed to be equal to the
strain in the concrete and can be used to estimate load using
the appropriate elastic modulus and section properties of the
shaft. A telltale is a metal rod installed within a hollow tube
embedded in the shaft. The bottom end of the rod is fixed at
a predetermined depth in the shaft and is the only point on the
rod in physical contact with the shaft. By measuring vertical
deformation of the upper end of the telltale during loading,
deformation of the shaft is determined for the depth at which
the telltale is fixed. By measuring the relative displacement
between two successive rods and distance between their
bottom ends, the average strain in the shaft between the two
telltales can be determined. Further information on these and
other types of instrumentation is given by Hirany and
Kulhawy (1988) and ONeill and Reese (1999).

The following case illustrates effective use of conventional axial load test on rock sockets. Zhan and Yin (2000)
describe axial load tests on two shafts for the purpose of confirming design allowable side and base resistance values in
moderately weathered volcanic rock for a Hong Kong transit project. The proposed design end bearing stress (7.5 MPa)
exceeded the value allowed by the Hong Kong Building
Code (5 MPa). One of the objectives of load testing was,
therefore, to demonstrate that a higher base resistance could
be used. The project involved 1,000 drilled shafts; therefore,
proving the higher proposed values offered considerable
potential cost savings.
Figure 58 shows the load test arrangement, consisting of
a loading platform for placement of dead load. Figure 59
shows details of one of the instrumented shafts. Strain gages
were provided at 17 different levels, including 4 levels of
gages in the rock socket. Two telltales were installed, one at
the base of the socket and one at the top of the socket. Shafts
were excavated through overburden soils using temporary
casing to the top of rock. When weathered rock was encountered, a 1.35-m-diameter reverse circulation drill (RCD) was
used to advance to the bearing rock, followed by a 1.05-mdiameter RCD to form the rock socket. For the shaft shown
in Figure 59, the socket was 2 m in length. A permanent,
bitumen-coated casing (to reduce side resistance in the overburden materials) was placed to the top of the socket. The
bottom was cleaned by airlift and concrete placed by tremie
(wet pour).

FIGURE 58 Axial load testing setup (Zhan and Yin 2000).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

77

FIGURE 60 Unit side and base resistance versus axial load


(Zhan and Yin 2000).

Testing Systems was undertaken to evaluate these and other


methods for deep foundations and to recommend interim
procedures for their use and interpretation. A draft final
report by Paikowsky et al. (2004b) describes these methods
in detail. The role of each of these tests for rock-socketed
shafts is described here.
FIGURE 59 Details of instrumented rock-socketed shaft (Zhan
and Yin 2000).

Figure 60 shows the results in terms of mobilized unit side


and base resistances versus load applied at the head of the
shaft. Unit side resistance reached a value of 2.63 MPa, well
exceeding the proposed design allowable value of 0.75 MPa.
Zhan and Yin noted that this value agrees well with Eq. 30 in
chapter three. Load transfer to the base was mobilized immediately upon loading, indicating excellent base conditions,
and reached a value exceeding 10 MPa. In the other shaft (not
shown) a unit base resistance of 20.8 MPa was reached with
no sign of approaching failure.

Osterberg Load Cell


The O-cell is a hydraulically operated jacking device that can
be embedded in a drilled shaft by attachment to the reinforcing cage (Figure 61). After concrete placement and curing, a
load test is conducted by expanding the cell against the portions of the shaft above and below it (Osterberg 1995). The
load is applied through hydraulic piston-type jacks acting
against the top and bottom cylindrical plates of the cell. The

The case presented by Zhan and Yin demonstrates how a


set of well-instrumented conventional axial load tests can be
used to (1) achieve cost savings on a project with a large
number of shafts, (2) confirm design allowable values of
socket resistance, (3) demonstrate suitability of the construction method, and (4) provide data against which design methods can be evaluated.
Conventional axial load testing has largely been replaced
by methods that are easier to set up and conduct, require less
equipment and space, are safer, less time consuming, and
usually less expensive, especially in rock. These methods
include the O-cell, Statnamic (STN), and dynamic impact
load tests. NCHRP Project 21-08, entitled Innovative Load

FIGURE 61 O-cell at bottom of reinforcing cage ready for


placement in a drilled shaft. (ONeill and Reese 1999).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

78

maximum test load is limited to the ultimate capacity of


either the section of shaft below the cell, the section above
the cell, or the capacity of the cell.
Pressure transducers are used to monitor hydraulic jack
pressures and converted to load. Linear vibrating wire displacement transducers (LVWDTs) between the two plates
measure total expansion of the cell and telltales are installed
to measure vertical movements at the top and bottom of the
test sections. The downward movement of the bottom plate
is obtained by subtracting the upward movement of the top
test section from the total extension of the O-cell as determined by the LVWDTs. Telltale deformations are monitored
with digital gages mounted on a reference beam. All of the
instrumentation is electronic and readings are collected by a
data acquisition unit.
The O-cell testing method provides some important advantages. There is no structural loading system at the ground
surface. Load can be applied at or very close to the base of a
socket for measurement of base resistance. In conventional
top load testing, most or all of the side resistance must be mobilized before there is significant load transfer to the base.
Some of the cited disadvantages are that the O-cell is sacrificial and requires prior installation, so it is not useful for testing existing foundations. Using a single O-cell, it is possible
to mobilize the ultimate capacity of one portion of the shaft
only, so that other sections of the shaft are not loaded to their
ultimate capacity.
According to DiMillio (1998), the majority of load tests
on drilled shafts are now being done with the O-cell. This is
supported by results of this study, in which 17 of 32 states responding to the survey reported using the O-cell for axial
load testing of rock-socketed shafts. Of these, 13 stated that
ultimate side resistance was determined and 7 reported that
the ultimate base resistance was determined. Five states indicated the test was used for proof load testing, in which design values of shaft resistance were verified. These responses
show that the O-cell has become a widely used method for
axial load testing of rock sockets.
A set of O-cell tests reported by Gunnink and Kiehne
(2002) serves to illustrate the type of information that is obtained from a typical test in which a single O-cell is installed
at the base of a rock socket. Figure 62 shows the test setup
for three test shafts socketed into Burlington limestone. As
shown, the shafts extended through soil before being socketed into limestone. All shafts were 0.46 m in diameter and
socket lengths ranged from 3.45 m to 3.85 m. Depth of soil
was approximately 4 m. Figure 63 shows test results for two
of the shafts (Shaft Nos. 1 and 3), respectively. Each graph
shows two curves, one of the O-cell load versus average measured uplift of the upper portion of the shaft, and the other of
the O-cell load versus downward displacement of the base of
the cell. Both figures are typical of failure of the shaft in uplift. At the maximum test load, it was not possible to main-

hydraulic lines

placement channel

dial gages

Test shaft
SOIL

ROCK

O-cell

FIGURE 62 Shaft and O-cell test setup (adapted from


Gunnink and Kiehne 2002).

tain or increase load without continuous upward deflection


of the top of the shaft, whereas the average base displacement
did not change. From these tests, it is not possible to determine ultimate base resistance values. The base load displacement curves show an interesting difference. For Shaft
No. 1, the downward base movement is small (around 1 mm)
up to the maximum test load, suggesting a very stiff base and
good contact between the concrete and underlying rock.
However, the curve for Shaft No. 3 shows downward movement approaching 10 mm upon application of the load, followed by a flattening of the curve. This behavior suggests the
presence of a compressible layer between the concrete and
underlying rock, possibly the result of inadequate cleanout of
the hole before pouring concrete. Both shafts were poured
under dry conditions and both were cleaned using the same
method, reported as rapidly spinning the auger bit after the
addition of water and then lifting out the rock cuttings.
Gunnink and Kiehne (2002) reported that it is common
practice to design drilled shafts founded in sound Burlington
limestone for base resistance only, using a presumptive allowable unit base resistance of 1.9 MPa. Side resistance is often neglected for design. Even the lowest observed base resistance measured by the O-cell tests yielded an allowable
unit base resistance of 5 MPa, assuming a factor of safety

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

79
30

30
upward displacement
of shaft above O-cell

20

Average displacement (mm)

Average Displacement (mm)

25

15
10
5
0
-5

downward displacement
of O-cell base plate

-10
-15

25
upward displ of shaft above O-cell

20
15
10
5
0
-5
-10

downward displ of O-cell base plate

-15
-20

-20
0

500

1000

1500

2000

2500

3000

3500

4000

500

1000

1500

2000

2500

3000

3500

4000

Load (kN)

Load (kN)

FIGURE 63 Results of single O-cell load tests: (left) Shaft No. 1; (right) Shaft No. 3 (Gunnink and Kiehne 2002).

of 3. The tests shown in Figure 63 yield ultimate unit side


resistances of 2.34 and 2.28 MPa, respectively. These tests
illustrate a typical outcome when field load testing is conducted; that is, measured unit side and base resistances
exceed presumptive values, sometimes significantly. Load

testing results make it possible to achieve more economical


designs. The O-cell tests also identify construction deficiencies, such as inadequate base cleanout (Figure 63 left).
The tests reported by Gunnink and Kiehne also illustrate
a limitation of testing with a single O-cell at the bottom of
the socket. The values of ultimate unit side resistance reported by the authors are based on the assumption that all
of the load was resisted by the rock socket, neglecting any
contribution of the overlying soil. It is not known how significant the error is for this case, but testing with multiple
O-cells makes it possible to isolate the section of shaft in rock
for evaluation of average side resistance (however, multiple
O-cells increase the cost of load testing). For example, if
a second O-cell is located at the top of the rock socket, a test
conducted with that cell can be used to determine the combined side resistance of all layers above the rock. An innovative approach based on this concept is illustrated in the testing
sequence shown in Figure 64. The figure and description are
from ONeill et al. (1997) based on tests conducted by
LOADTEST, Inc., for the Alabama DOT. Arrangement of the
O-cells and the 4-step testing sequence depicted in the figure
made it possible to measure ultimate base resistance, side resistance of the socket (in both directions), and side resistance
of the cased portion of the shaft above the socket. It is noted
that this arrangement made it possible to measure a total
foundation resistance of 80 MN, compared with approximately 11 MN for the largest standard surface jacks. Installation of multiple O-cells makes it necessary to provide a
tremie bypass line to facilitate placement of concrete below
and around the upper cells.
Interpretation of O-cell tests in rock sockets is typically
based on the assumption that total applied load at the ultimate
condition is distributed uniformly over the shaft/rock side
interface, and used to calculate an average unit side resistance by

FIGURE 64 Test setup and loading sequence with two O-cells


(ONeill et al. 1997).

fs =

Qoc
BD

Copyright National Academy of Sciences. All rights reserved.

(148)

Rock-Socketed Shafts for Highway Structure Foundations

80

where
fs = average unit side resistance (stress),
QOC = O-cell test load,
B = shaft diameter, and
D = socket length.
The degree to which this average unit side resistance is valid
for design of rock sockets loaded at the head depends on the
degree to which side load transfer under O-cell test conditions is similar to conditions under head loading. Detailed
knowledge of site stratigraphy is needed to interpret side load
transfer.
O-cell test results typically are used to construct an equivalent top-loaded settlement curve, as illustrated in Figure 65.
At equivalent values of displacement both components of
load are added. For example, in Figure 65a, the displacement
for both points labeled 4 is 10 mm. The measured upward
and downward loads determined for this displacement are
80
12

60
11
10

Movement (mm)

40

side load-deformation
curve is measured

20

6 7

side resistance curve


is extrapolated

0
1

-20

7
8

-40

9
8

measured base loaddeformation curve

9
10
11

-60

12

-80
0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

16.0

O-Cell Load (MN)

(a)

Equivalent Top Load (MN)

30

25

20
6
5

12

11

10

7
base resistance measured,
side resistance extrapolated

15
4
3

10

0
0

10

20

30

40

50

60

70

Downward Displacement (mm)


(b)

FIGURE 65 Construction of equivalent top-loaded settlement


curve from O-cell test results (a) O-cell measured loaddisplacement; (b) equivalent top-load settlement results.

80

added to obtain the equivalent top load for a downward displacement of 10 mm and plotted on a load-displacement
curve as shown in Figure 65b. This procedure is used to
obtain points on the load-displacement curve up to a displacement corresponding to the least of the two values (side or
base displacement) at the maximum test load. In Figure 65a,
this corresponds to side displacement. Total resistance corresponding to further displacements is approximated as follows. For the section of shaft loaded to higher displacement,
the actual measured load can be determined for each value of
displacement up to the maximum test load (in Figure 65a this
is the base resistance curve). The resistance provided by the
other section must be estimated by extrapolating its curve
beyond the maximum test load. In Figure 65a, the side resistance curve is extrapolated. The resulting equivalent toploaded settlement curve shown in Figure 65b is therefore
based on direct measurements up to a certain point, and partially on extrapolated estimates beyond that point.
According to Paikowsky et al. (2004b), most state DOT
geotechnical engineers using O-cell testing tend to accept
the measurements as indicative of drilled shaft performance
under conventional top-down loading. O-cell test results are
applied in design by construction of an equivalent top-load
settlement curve, as illustrated earlier, or by using the
measured unit side and base resistances as design nominal
values. However, some researchers (ONeill et al. 1997;
Paikowsky et al. 2004b) have pointed out differences between O-cell test conditions and top loading conditions that
may require interpretation. The most significant difference is
that compressional loading at the head of a shaft causes compression in the concrete, outward radial strain (Poissons
effect), and a load transfer distribution in which axial load in
the shaft decreases with depth. Loading from an embedded
O-cell also produces compression in the concrete, but a load
transfer distribution in which axial load in the shaft decreases
upward from a maximum at the O-cell to zero at the head of
the shaft. It is possible that different load transfer distributions could result in different distributions of side resistance
with depth and, depending on subsurface conditions, different total side resistance of a rock socket.
In shallow rock sockets under bottom-up (O-cell) loading
conditions, a potential failure mode is by formation of a conical wedge-type failure surface (cone breakout). This type
of failure mode would not yield results equivalent to a shaft
loaded in compression from the top. A construction detail
noted by Crapps and Schmertmann (2002) that could potentially influence load test results is the change in shaft diameter that might exist at the top of a rock socket. A common
practice is to use temporary casing to the top of rock, followed by a change in the tooling and a decrease in the diameter of the rock socket relative to the diameter of the shaft
above the socket. Top-down compression loading produces
perimeter bearing stress at the diameter change as illustrated
in Figure 66, whereas loading from an O-cell at the bottom
of the socket would lift the shaft from the bearing surface.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

81

compression

SOIL
change in diameter at
soil/rock interface

ROCK
FIGURE 67 Comparison of load-displacement curves; O-cell
versus FEM (Paikowsky et al. 2004b).
FIGURE 66 Perimeter bearing stress at diameter change
under top loading.

Paikowsky et al. (2004b) reviewed the available data that


might allow direct comparisons between O-cell and conventional top-down loading tests on drilled shafts. Three sets of
load tests reported in the literature and involving rock sockets
were reviewed. However, in two of the cases the test sequence
involved conventional top-down compression loading (Phase
1) followed by O-cell testing from the bottom up (Phase 2).
Mobilization of side resistance in Phase 1 is believed to have
caused a loss of bond, thereby influencing results of the O-cell
tests and precluding any direct comparison. The third case involved STN and O-cell tests of shafts in Florida limestone.
Paikowsky et al. stated that several factors, including highly
variable site conditions and factors related to the tests, prevented a direct comparison of results.
FEM reported by Paikowsky et al. (2004b) suggests that
differences in rock-socket response between O-cell testing
and top-load testing may be affected by (1) modulus of the
rock mass, EM, and (2) interface friction angle, i. Paikowsky
first calibrated the FEM model to provide good agreement
with the results of O-cell tests on full-scale rock-socketed
shafts, including a test shaft socketed into shale in Wilsonville,
Alabama, and a test shaft in claystone in Denver, Colorado,
described by Abu-Hejleh et al. (2003). In the FEM, load was
applied similarly to the field O-cell test; that is, loading from
the bottom upward. The model was then used to predict behavior of the test shafts under a compression load applied at
the top and compared with the equivalent top-load settlement
curve determined from O-cell test results. Figure 67 shows a
comparison of the top-load versus displacement curves for
the Alabama test, one as calculated from the O-cell test and
the other as predicted by FEM analysis. The curves show
good agreement at small displacement (<0.1 in. or 2.5 mm);
however, the curve derived from FEM analysis is much
stiffer at higher displacement. This exercise suggests that the

equivalent top-load settlement curve derived from an O-cell


load test may underpredict side resistance for higher displacements; that is, the O-cell derived curve is conservative.
Further FEM analyses reported by Paikowsky et al. (2004b)
suggest that the differences between loading from the bottom (O-cell) and loading in compression from the top are the
result of differing normal stress conditions at the interface,
and that these differences become more significant with increasing rock mass modulus and increasing interface friction
angle.
These numerical analyses suggest that differences in the
response of rock sockets to O-cell test loading and top-down
compression loading may warrant consideration in some
cases. Ideally, side-by-side comparisons on identical test
shafts constructed in the same manner and in rock with similar characteristics and properties are needed to assess differences in response. However, it is expected that the potential
differences, if any, will eventually be identified and incorporated into interpretation methods for O-cell testing. In the
meantime, the O-cell test is providing state transportation
agencies with a practical and cost-effective tool for evaluating the performance of rock sockets and it is expected that
the O-cell test will continue to be used extensively.
Instrumentation such as sister bars with strain gages
makes it possible to better determine the load distribution and
load transfer behavior during an O-cell load test. This information can then be used to make more refined predictions of
load transfer behavior under head load conditions.
In summary, some of the advantages of the O-cell for axial load testing of rock-socketed shafts include:
Ability to apply larger loads than any of the available
methods (important for rock sockets) and

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

82

With multiple cells or proper instrumentation, it can


isolate socket base and side resistances from resistance
of other geomaterial layers.
Limitations of the O-cell test for use by state DOTs include:
Shaft to be tested must be predetermined, because it is
not possible to test an existing shaft;
For each installed device, test is limited to failure of one
part of the shaft only;
There are possible concerns using test shaft as a production shaft;
Interpretation methods that account for differences in
loading mode are not yet fully developed; and
There are currently no ASTM or AASHTO standards
specifically for O-cell load tests.
Interviews with state DOT engineers for this study show
that the O-cell test has been an integral tool in advancing the
understanding and use of rock-socketed drilled shafts. The
Kansas DOT (KDOT) experience is representative of several
other states. The following is based on an interview with
Robert Henthorne, KDOT Chief Geologist. The geology
of the western half of Kansas, located in the High Plains
physiographic province, is dominated by thick sequences of
sedimentary rocks, mostly sandstone, shale, and limestone.
Until approximately 1995, virtually all highway bridges were
founded on shallow foundations or H-piles driven to refusal
on rock. Drilled shafts were not considered a viable alternative because of uncertainties associated with both design and
construction. With encouragement from FHWA, KDOT
engineers and geologists initiated a long-term program of
training, education, and field load testing to better match
foundation technologies with subsurface conditions. Workshops on drilled shaft design, construction, inspection, and
nondestructive testing (NDT), sponsored by FHWA and the
International Association of Foundation Drilling (ADSC),
were conducted at the invitation of KDOT. KDOT began using drilled shafts as bridge foundations where appropriate.
Several bridge sites in western Kansas were designed with
rock-socketed shafts. To address the lack of experience with
these conditions, O-cell testing was incorporated into the larger
bridge projects. In almost every case, the O-cell test results
showed side and base resistances considerably higher than
the values used for preliminary sizing of the shafts, and valuable
experience was gained with construction methods, effective
cleanout strategies, NDT methods, etc. KDOT now has O-cell
test results on rock-socketed shafts from nine projects and has
developed in-house correlations between rock mass properties
and design parameters for commonly encountered geological
formations. Drilled shafts now comprise approximately 70%
to 80% of new bridge foundations, and shaft designs are
more economical because there is a high level of confidence in
capacity predictions, based directly on the load tests.
The approach taken by KDOT illustrates how field load
testing, in this case with the O-cell, can be incorporated into

an overall program leading to increased use and improved design methods for rock-socketed foundations. The Colorado
DOT has also used O-cell testing to improve its design
procedures for rock-socketed shafts, as documented by AbuHejleh et al. (2003).
Statnamic
The STN load test was developed in the late 1980s by
Berminghammer Foundation Equipment of Hamilton, Ontario. Its use in the U.S. transportation industry has been
supported by FHWA through sponsorship of load testing
programs, as well as tests conducted with an STN device
owned by FHWA for research purposes.
In this test, load is applied to the top of a deep foundation
by igniting a high-energy, fast-burning solid fuel within a
pressure chamber. As the fuel pressure increases, a set of reaction masses is accelerated upward, generating a downward
force on the foundation element equal to the product of the
reaction mass and the acceleration. Loading occurs over a
period of approximately 100 to 200 ms, followed by venting
of the pressure to control the unloading cycle. Load applied
to the foundation is monitored by a load cell and displacement is monitored with a photovoltaic laser sensor. The concept is illustrated schematically in Figure 68. STN equipment
is available for test loads as high as 30 MN.
Processing of the load and displacement time histories is
required to convert the STN measurements into an equivalent,
static load-displacement curve. The analysis accounts for
dynamic effects that may include damping and inertial
effects. The unloading-point method as reported by Horvath
et al. (1993) provides a relatively straightforward method for
determining static resistance using measurements made at the
top of the shaft during a STN test. Test interpretation is also
discussed by Brown (1994) and El Naggar and Baldinelli
(2000).

FIGURE 68 Schematic of STN load test (ONeill et al. 1997).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

83

Mullins et al. (2002) recently introduced the segmental


unloading point method, which uses top and toe measurements as well as strain measurements from along the length
of the foundation. The segmental unloading point method
enables determination of load transfer along various segments of the foundation, an advantage for rock-socketed
shafts to separate resistance developed in rock from that
developed in portions of the shaft embedded in soil. The analysis is automated using software provided by the testing firm
and equivalent static load-displacement graphs are produced
immediately for evaluation. All data are stored for future
analysis and reference.
During the 1990s, FHWA performed or funded STN and
correlation studies with conventional static load tests to
develop standardized testing procedures and data interpretation methods (Bermingham et al. 1994). Numerous other
studies have further expanded the database of case histories
and performance studies. The result is that STN testing is
now a well-developed technology that is highly suitable for
use by state DOTs for axial load testing of rock-socketed
shafts. STN advantages identified by Brown (2000) include:

Large load capacity, applied at top of shaft;


Can test existing or production shaft;
Economies of scale for multiple tests;
Amenable to verification testing on production shafts;
and
Reaction system not needed.
Disadvantages include:
Capacity high, but still limited (30 MN);
Rapid loading method, as rate effects can be significant
in some soils (less in rock);
Mobilization costs for reaction weights; and
Not currently addressed by ASTM or AASHTO standards.
Mullins, as reported in Paikowsky et al. (2004b), analyzed
a database of 34 sites at which both STN and static load tests
were conducted on deep foundations. The data included load
tests on four drilled shafts in rock at two sites, one site each
in Florida and Taiwan. The objective of the study was to
develop recommendations for LRFD resistance factors when
axial compression capacity is based on STN testing. The authors recommend a resistance factor of 0.74 for all deep foundation types in rock (not specific to drilled shafts) when
tested by STN. In addition, a rate effect factor (REF) is recommended to account for rate effects when using STN
results by the unloading point method. The REF varies with
soil or rock type and recommendations are given here. If the
segmental unloading point method is used (requiring strain
gages), separate REF factors can be applied to each segment to account for different soil or rock types. This analysis addresses the disadvantage cited previously regarding rate
effects.

Derived Static = REF*UP-derived capacity


REF = 0.96 for rock
0.91 for sand
0.69 for silt
0.65 for clay.

Dynamic Impact Testing


A dynamic compression load test can be carried out by dropping a heavy weight onto the head of the shaft from various
heights. The shaft is instrumented with strain gages and
accelerometers to measure the force and impact velocity of
the stress wave generated by the dynamic impact. The measurements are correlated to driving resistance to predict load
capacity. A review of various available drop weight systems
and evaluation of the method is given by Paikowsky et al.
(2004c). A typical drop weight system consists of four
components: (1) a frame or guide for the drop weight, (2) the
drop weight (ram), (3) a trip mechanism to release the ram, and
(4) a striker plate or cushion, as shown in Figure 69. Various
configurations of modular weights can be used to provide
ram weights as high as 265 kN (Hussein et al. 2004) and drop
heights are adjustable up to 5 m (Paikowsky et al. 2004c). A
rule of thumb given by Hussein et al. is that a ram weight of
1% to 2% of the expected shaft capacity be available on site.
Drop weight load testing interpretation relies on analysis
methods similar to those used in standard dynamic pile testing. Strain gage and accelerometer measurements at the top
of the pile are used to evaluate characteristics of stress wave
propagation. If sufficient shaft resistance is mobilized, it is
possible in theory to relate the stress wave characteristics to
shaft capacity using available PDA (Pile Driving Analyzer)
technologies. Drop weight testing of drilled shafts has not
been used extensively on bridge foundations in the United
States, in part because other available methods (e.g., O-cell
and STN) provide a more direct measurement of static resistance. According to DiMillio (1998), test results on FHWA
projects have not demonstrated sufficiently good agreement
between drop weight and other tests. The drop weight tests
reportedly overpredicted measured capacities.
Drop weight testing for rock sockets is suitable for postconstruction tests at bridge sites where questions arise during
construction regarding the performance of as-built foundations. This application is illustrated by the case of the Lee
Roy Selmon Crosstown Expressway, in Tampa, Florida.
The columns supporting an elevated section of roadway are
founded on drilled shafts socketed into limestone. During
construction of the superstructure, one of the columns suddenly underwent more than 3 m (11 ft) of settlement as a
result of the failure of the drilled shafts. Subsequent investigations determined that the failed shafts were not founded
in sound limestone as believed, raising questions about the
capacity of all 218 drilled shafts supporting the elevated
roadway. As part of an investigation to determine how many

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

84

FIGURE 69 Schematic of drop weight system (Paikowsky et al. 2004c).

shafts might need remediation, dynamic load tests were conducted on 12 of the shafts supporting existing columns using
the pile driving hammer shown in Figure 70. Testing proved
the design capacity of 11 of the 12 shafts tested. This case
also illustrates the need for thorough subsurface investigation when socketing into limestone. In this case, rock elevations were found to be highly variable. Seismic methods
used in combination with borings in the post-failure investigation provided a more detailed geologic model of site
conditions.

Interpretation Framework for Static Axial


Load Tests

Carter and Kulhawy (1988) and Kulhawy and Carter (1992b)


proposed a method for interpretation of static axial load tests
on rock-socketed shafts. The method involves analyzing a
static axial load-displacement curve from a load test according to the analytical closed-form solutions presented in chapter three (Eqs. 6995). The parameters back-calculated from
the load test could then be used to evaluate effects of various
design parameters on the load-displacement behavior of trial
designs that differ from that of the test shaft. The method is

applicable to shafts that satisfy the criteria for rigid behavior,


given as
2

Ec
Er 1
2D
B

(149)

in which Ec = modulus of the reinforced-concrete shaft, Er =


rock mass modulus, D = socket length, and B = socket
diameter.
The analysis is applied to two cases: (1) shear socket under compression or uplift and (2) complete socket under
compression. The shape of a load-displacement curve from a
load test is modeled in terms of constant slopes (S), which are
related mathematically to the model parameters described in
chapter three. Consider the load-displacement curve for a
shear socket loaded in compression, as shown in Figure 71.
Three parameters are required to idealize the geometry of
the curve. S1 is the slope of the initial portion, S2 is the
approximated slope of the full-slip portion of the curve, and
Qi is the intercept on the vertical axis (wc = 0) of the line with
slope S2. For a rigid shaft, the measured curve parameters

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

85

theoretically are related to the elastic model parameters by


the relationships given here (a):
(a) Shear Socket Compression or Uplift
(1 + r )
Er =
S1
D
1 S2
tan tan =
2 S1 S2
Q
c = ( 2 tan tan + 1) i
BD
with
D
= ln 5 (1 r )
B

(150)

in which r = Poissons ratio of the rock mass, = interface


friction angle, and = interface angle of dilation, and c = interface cohesion.
For a complete socket under compression in which the base
load-displacement is determined (Figure 72), the loaddisplacement curve is approximated by S1, S2, Qi, and S3, the
slope of the base load-displacement curve. The curve parameters are related to the elastic model parameters as given in (b),
(b) Complete Socket Compression

FIGURE 70 Dynamic load testing of shaft-supported column in


Tampa, Florida.

(1 + r )
Er =
( S1 S3 )
D
1 + b 2
Eb =
S3
B
1 S S3
tan tan = 2
2 S1 S2
Q
c = ( 2 tan tan + 1) i
BD

in which Eb = modulus of the rock mass beneath the shaft base.


Carter and Kulhawy (1988) applied the technique described
to 25 axial load tests reported in the literature by backcalculating values of the model parameters Er, Eb, c, and (tan
tan) from load-displacement curves using the equations
given previously. A limitation of the model described earlier
is that the assumption of rigidity may be less acceptable for
shafts in harder rocks where the modulus values for the rock
mass and the shaft material are closer. The reader is advised to
review the original publications for further assumptions and
derivations of the equations.
Lateral Load Testing
FIGURE 71 Interpretation of a side-shear-only test (Carter and
Kulhawy 1988).

A significant number of states indicated in the questionnaire


that lateral loading governs the design of rock-socketed

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

86

depths corresponding to approximately the first 10 diameters


of the shaft. Other important points to consider when conducting conventional lateral load tests, as pointed out by
ONeill and Reese (1999) are summarized as follows.
The test site conditions and test shaft should be selected
and built to match as closely as possible the actual conditions
to which they will be applied. Items such as overburden
stresses acting in the resisting soil and rock layers, groundwater and surface water conditions, shaft dimensions and reinforcing, and construction methods all can have a significant
influence on the lateral load response of a drilled shaft. To
the extent possible, these conditions should be matched by
those of the load test.

FIGURE 72 Interpretation of a complete socket test (Carter and


Kulhawy 1988).

shafts for a significant percentage of projects (Question 25).


However, as noted in chapter four, very few lateral load tests
have been conducted on rock-socketed shafts. Methods for
conducting lateral load tests on deep foundations include
conventional methods, Osterberg load cell, and STN.
Conventional Lateral Load Test
The conventional method for conducting a lateral load test is
given in ASTM D3966 and involves pushing or pulling the
head of the test shaft against one or more reaction piles or
shafts. A variety of arrangements for the test shaft and reaction shaft are possible and these are given in detail in Reese
(1984) and Hirany and Kulhawy (1988). One approach is to
use two shafts and apply the load such that both shafts are
tested simultaneously, providing a comparison between two
shafts. A load cell is used to measure the applied lateral load
and dial gages or displacement transducers attached to a reference beam can be used to monitor lateral deformation.
Thorough treatment of instrumentation for lateral load tests
can be found in Reese (1984) and Hirany and Kulhawy
(1988).
Drilled shafts are often used where the designer wishes to
take advantage of their large lateral load capacity, especially
that of large-diameter shafts. Analysis often shows that the
geomaterials in the upper part of the ground profile have the
most significant influence on lateral deformations and lateral
load transfer. A critical part of lateral load testing is to have
detailed knowledge of the site stratigraphy, particularly at the

Analysis of the load test results will be interpreted using


the analytical methods presented in chapter four. The most
widely used method is the p-y curve method, in which p-y
curves are fit to obtain agreement with the load test measurements. As a minimum, it is therefore necessary to have
reliable measurements of ground line shear load, ground line
deflection, and rotation (requires two deflection points separated by a known vertical distance). To define p-y curves accurately over the length of the shaft requires measurements
of the deflected shape of the shaft, which can be done using
slope inclinometer measurements. A more accurate method
to determine p-y curves (or to evaluate any analytical
method) is to establish bending moment as a function of
depth, which can be done by installing a steel tube with
closely spaced strain gages along the length of the shaft. This
approach is most appropriate for tests conducted for applied
research; for example, to develop new methods for establishing p-y curves in rock.
Boundary conditions must be considered carefully when
back-fitting analytical models and then applying the model
for design. In a lateral load test, the boundary conditions at
the head of the shaft will normally be free of any rotational
restraint and have zero applied moment and zero axial load.
Service boundary head conditions are likely to include some
head restraint and possibly axial load and moment. Also, the
nonlinear momentEI relationships must be accounted for
both in the load test and in the analysis.
Four states (California, Massachusetts, New Jersey, and
North Carolina) reported the use of conventional lateral load
tests on rock-socketed shafts. Although lateral load testing is
not as common as axial testing, conventional testing has been
the method of choice for lateral. Other methods have, so far,
been used on a limited basis. These include lateral O-cell
tests (at least two states, South Carolina and Minnesota) and
lateral STN (Alabama, Florida, Kentucky, North Carolina,
South Carolina, and Utah). Several states that did not respond
to the survey are known to have conducted lateral STN (Ohio
and Virginia).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

87

Lateral Osterberg Load Test


The O-cell can be embedded in a drilled shaft and oriented
such that the load is applied in the horizontal direction. The
method is described by ONeill et al. (1997) for a case in
which the Minnesota DOT required representative p-y curves
for a stratum of friable sandstone situated beneath a thick
layer of normally consolidated clay. Shafts socketed into the
sandstone were to support a bridge undergoing ice loading.
The test was conducted at a nearby location in which a 26.7
MN O-cell was positioned vertically within a 1.22-m-diameter
socket, as depicted in Figure 73, and used to thrust the two
halves of the socket against the rock. Lateral force and deflection measurements were used to derive p-y curves. The
authors point out that care must be taken in interpreting
the results, because the stressstrain conditions created by
the test are not the same as in a laterally loaded socketed shaft
that is loaded at its head and not split.
Lateral O-cell testing of rock sockets offers some of the
same advantages as for axial O-cell load testing, namely the
elimination of a structural loading system at the ground level.
Also, the test provides the ability to apply lateral loading at predetermined depths, such as within the rock socket. Further research is needed to establish guidelines for proper procedures
and to define correct analyses that account for the differences
in boundary conditions, load transfer, and soil and rock resistance, compared with a shaft loaded at its head. It is also worth
noting that the lateral split socket test may provide a means to
measure the in situ rock mass modulus of deformation (EM).
Lateral Statnamic
The STN load test has also been adapted for lateral loading.
The device is mounted on steel skids supported on the ground

FIGURE 73 Top view of O-cell arrangement for lateral split


socket test (ONeill 1997).

allowing the reaction masses to slide on rails, as shown in


Figure 74. The lateral STN test can simulate lateral impact
loading such as might occur against a bridge pier from a
vessel.
The lateral STN test can also be used to derive the static
lateral response, but requires appropriate instrumentation
and correct analysis of the test results. In tests described by
Brown (2000), the following instrumentation was used:

Load cell,
Displacement transducers,
Accelerometers on top of cap or shaft,
Downhole motion sensors,
Resistance-type strain gages, and
Megadac Data Acquisition System.

Figure 75a shows the measured dynamic response of the


shaft in terms of force, acceleration, and lateral displacements
versus time. The curves showing measured lateral displacement from three measurements are identical and cannot be
distinguished in the figure. Dynamic response is separated
into static, inertial, and damping components. A p-y analysis
(using LPILE or FBPIER) is fit to obtain a reasonable match
between the measured load-displacement response for each
component of force (static, inertial, and damping). Load versus displacement curves derived are shown in Figure 75b
based on analysis of the dynamic response in Figure 75a.
The lateral STN test is reported as to be safe, controlled,
and economical. Its principal advantage lies in the ability to
measure directly the dynamic lateral response and to provide
a derived static response. This test is a valuable tool for the
design of bridge foundations to withstand dynamic lateral
loading from earthquakes, wind, and vessel impacts. The test
may also be used in place of a conventional static test. Lateral
loads up to 18 MN may be possible.

FIGURE 74 Lateral STN load test (Courtesy: L. Fontaine).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

88
Shaft Group, Load 5 (E-W)
1000
800

1.5

600

400

0.5

200
0

0
-0.5

-200

-1

-400

Load, tons
Statnamic Acceleration
tons

Load

inches / g's

Displacement / Acceleration

2.5

Lateral Translation, Top West


Lateral Translation, Top East
Lateral Translation, Top Center

-600

-1.5

-800

-2
0.4

0.6

0.8

1.2

seconds

Time

(a)
Lateral Load versus Translation
140

600

120

500

100

400

80

300

60

200

40

100

20

% Damping

tons

Load

Static & Derived Statnamic - Shaft Group


700

Static
Derived Statnamic
% Damping
Total Resistance (Static + Damping)

0
0

0.5

1.5

2.5

inches

Tlti

(b)

FIGURE 75 Results of lateral STN test: (a) measured dynamic


response; (b) derived static response.

CONSTRUCTABILITY, INSPECTION, AND


QUALITY ASSURANCE

These topics are considered together because they encompass activities having a single objective: construction of
a high-quality, rock-socketed drilled shaft foundation that
performs in accordance with the design assumptions. As
illustrated in the flow chart diagram of Figure 3, chapter one,
the final design is based on input from three general sources:
(1) site characterization, (2) geotechnical analysis, and
(3) structural analysis and modeling. Plans and specifications
are developed that reflect generally accepted practices based
on the collective experience of the construction and engineering communities. Examples of model specifications
include those given in Chapter 15 of the FHWA Drilled Shaft
Manual (ONeill and Reese 1999), ACI Standard Specification for the Construction of Drilled Piers, ACI 336.1-98
(1998), and specifications developed by state and federal
transportation agencies with extensive experience in drilled
shaft use. In addition, effective specifications will address
issues that are unique to the specific conditions that determine
the final design, including constructability issues which, ideally, are accounted for in all three of the input categories identified previously. In the following paragraphs, these topics are
discussed individually, but in practice they must be integrated
into the design concepts discussed in this synthesis.
Constructability

Much emphasis has been placed on constructability of drilled


shafts by FHWA and through efforts of the International As-

sociation of Foundation Drilling (ADSC). The FHWA


Drilled Shaft Manual (ONeill and Reese 1999) addresses
constructability and its role in drilled shaft design. The manual also forms the basis of a National Highway Institute (NHI)
course on drilled shafts that is available through FHWA. A
separate NHI course that certifies inspectors for drilled shaft
construction (Williams et al. 2002) also has a strong emphasis on constructability and was developed with significant
contractor input. ADSC provides short courses, workshops,
and a library of publications focused on construction-related
issues for drilled shafts. ADSC also provides constructability reviews of individual projects in which independent contractors review the project plans and specifications and offer
advice on its constructability. This step could be incorporated
into the overall process depicted in Figure 3, as denoted in the
flow chart by constructability review.
Integrating constructability into a drilled shaft project
involves taking a common sense approach to design that
accounts for the methods, tools, and equipment used by contractors to build the shafts. No attempt will be made here to
identify all of these issues, but items identified by the survey
and that relate specifically to rock sockets are discussed.
Schmertmann et al. (1998) and Brown (2004) present
guidelines for ensuring quality in drilled shaft construction
and some recent advances in materials that have applications
in both soil and rock. The key elements to be considered to
avoid the most commonly observed construction problems
are:

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

89

Workability of concrete for the duration of the pour;


Compatibility of congested rebar and concrete;
Control of stability of the hole during excavation and
concrete placement, especially with casing;
Proper consideration and control of hydrostatic balance
and seepage;
Bottom cleaning techniques and inspection; and
Drilling fluid that avoids contamination of the bond
between concrete and bearing material or excessive
suspended sediment.
New developments in concrete mix design, in particular
mixes described as self-consolidating concrete (SCC), can
provide benefits for drilled shaft construction. The characteristic of SCC that is most beneficial is very high slump
flow. Reinforcement cages with a high density of steel bars,
often necessary especially for seismic design, make it difficult to provide the necessary clear spacing between bars that
will ensure flow of concrete to the outside of the cage. The
flow properties of SCC have been shown to reduce potential
defects associated with incomplete cover or voids caused by
inadequate flow of concrete.
Prompt placement of concrete is another construction
practice that promotes quality in the as-built shaft. Delay in
concrete placement increases the potential for slump loss
and, in some cases, has been identified as a cause of reduced
side resistance (Schmertmann et al. 1998).
Several states identified problematic construction issues
when the slurry method of construction is used in rock sockets. One issue is whether slurry has a detrimental effect on
side resistance of rock sockets. Thirteen states indicated that
they restrict the use of slurry in rock sockets and one state
expressed concerns with use of drilling fluids instead of casing. In many situations, if casing is used to support the hole,
the need to use slurry is eliminated. Typically, casing need
only extend to the top of rock if the rock-socket portion of
the hole will remain open without caving. If there is water
in the overburden, the casing can be sealed into the rock,
dewatered, and the socket can then be excavated without support. However, there are situations where a contractor may
deem it necessary to introduce slurry. For example, when
rock is highly fractured it may not be possible to seal the casing sufficiently to prevent water inflow, and a contractor may
elect to use slurry. In this case, slurry may be used to balance
the hydraulic head to prevent seepage into the hole that can
disturb the material at the base of the shaft, an issue related
directly to design decisions on whether to include base resistance in the design. For reverse circulation drilling, slurry
may be used as the circulating fluid (e.g., the RichmondSan
Rafael Bridge shown in Figure 56).
There are few data showing the effects of properly mixed
and handled slurry on rock-socket side or base resistances.
Slurry that does not possess the appropriate viscosity, density,
and sediment content, or that is allowed to remain in the hole
(and not agitated) long enough to form a thick filter cake, will

almost certainly reduce side resistance compared with a shaft


drilled and poured under dry conditions, in either soil or rock.
However, if sound practices are followed by an experienced
contractor and there is proper inspection, slurry drilling for
rock sockets can be an effective construction method, assuming the slurry is handled in a manner that avoids contamination of the interface bond or excessive suspended sediment.
In certain rock types, there is evidence that use of polymer
slurry may be beneficial to rock-socket side resistance. The
Kentucky DOT requires polymer slurry for drilling in rock
that exhibits low values of slake durability index. Typically,
this is the case in certain shale formations in Kentucky. Slaking occurs when the shale is exposed to water, and can cause
formation of a smear zone, reducing side resistance considerably, as demonstrated by Hassan and ONeill (1997).
Apparently, the polymer slurry prevents softening and the
resulting smear zone, although there have not been load tests
in which a direct comparison has been made. This issue
deserves further research.
One state DOT identified the following as a problematic
construction issue: various methods used to force a dry
pour, indicating that some measures taken to avoid placing
concrete under water or slurry are more detrimental than
allowing a wet pour. Both Schmertmann et al. (1998) and
Brown (2004) describe a case that seems to contradict some
commonly held ideas about casing versus wet hole construction of rock sockets. A drilled shaft installed through 12 m
of soil and socketed into rock was constructed using a fulllength casing (to provide downhole visual inspection). A
load test using the Osterberg load cell indicated a mobilized
side resistance in the socket of 0.5 MN, much less than
expected. A second shaft was constructed, but using a wet
hole method with tremie placement of concrete and without
casing into the rock. Load testing of this shaft indicated more
than 10 MN of side resistance in the socket. The difference
is attributed to a decrease in concrete workability during the
time required to remove the casing after concrete placement,
preventing formation of a good bond along the socket interface. Trapping of debris between the casing and rock could
also have occurred and may have smeared cuttings along the
sidewalls. The lesson of this case is that the construction
method should be selected to provide the best product for the
given conditions, and that in many situations a wet hole
method is the most effective and will not adversely affect
shaft behavior if done properly. Forcing a dry pour may
cause more problems than it solves.
Another good reason to review the ground conditions carefully before allowing dry hole construction is identified
by Schmertmann et al. (1998). If the groundwater elevation
is above the base of the hole, dry conditions inside the socket
result in a hydraulic gradient causing inward seepage as
illustrated in Figure 76. They describe several cases where
seepage degraded side resistance and base resistance. Maintaining a slurry or water level inside the hole sufficient to balance the groundwater pressure eliminates the inward gradient

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

90
casing

Zone of disturbed material at


base caused by inward seepage

FIGURE 76 Development of disturbed base caused by high


seepage gradient toward bottom of a cased hole.

and prevents base and side disturbance. The authors cite


several cases in which comparisons of Osterberg load cell test
results on shafts poured both wet and dry show this effect.
The most common factor cited in construction claims associated with rock-socketed shafts is differing site conditions,
that is, the subsurface conditions actually encountered during
construction are claimed to be materially different from those
shown in boring logs. Responses to the questionnaire did not
indicate that claims were a major obstacle to the use of drilled
shafts for most states. However, one state DOT gave the following response when asked to comment on issues pertaining to the use of rock-socketed drilled shafts by your agency
(Question 6): Most result in claims due to the requirement to
include Differing Site Conditions on all contracts.
The same agency responded as follows to Question 36
pertaining to perceptions of construction problems:
We design for low bidding contractors to get the contract and the
construction problems that will result. Rock may be harder than
the contractor thought when bidding and planning the job. Thus
the drilling equipment brought out is often unable to drill or very
slow to drill the rock. This results in costly contractor claims.

Claims for differing site conditions are part of the geotechnical construction field, but measures can be taken to
minimize them. For example, one contractor interviewed for
this study noted that geotechnical reports often place strong
emphasis on rock of the lowest strength, because these layers may control side or base resistances for design. However,
for estimating drilling costs, contractors need information on
rock layers of the highest strength, because that will dictate
the type of drilling and tools needed to bid the job accurately
and to carry out the construction properly. Transportation
agencies might consider surveying contractors to find out
exactly what information contained in their boring logs is
most helpful for bidding on rock-socket jobs, and what addi-

tional information could improve their ability to perform the


work. Another contractor interviewed for this study stated
that the rock classification system of the ISRM is useful to
determine what type of tool (rock auger, core barrel, or
downhole hammer) will be most effective. The ISRM system
places rock into one of seven categories (R0 through R6)
based on strength, as described in chapter two (see Rock
Material Descriptors).
An issue identified by several states is the discrepancy
that sometimes occurs between the elevation corresponding
to top of rock as shown in boring logs and as encountered
during construction. The Washington State DOT uses language in their special provision for rock sockets that reportedly works well and is summarized as follows. For shafts
with a specified minimum penetration into the bearing layer
and no specified base elevation, the contractor furnishes each
reinforcing cage 20% longer than specified in the plans. The
increased length is added to the bottom of the cage. The contractor then trims the reinforcing cage to the proper length
before placement. The DOT assumes the cost of the excess
steel, but believes that cost is offset by avoiding construction
delays, disputes, and claims that may occur otherwise.
Other specific issues identified by states in the questionnaire
pertain to inadequate cleanout buckets, improper placement
of concrete with pump trucks, and a case in which temporary
casing to support the overburden with the same diameter as the
rock socket resulted in the casing being dragged down into
the socket, requiring additional socket drilling. There is a
constructability lesson in each of these cases.
Certain geologic conditions are associated with more
challenging construction and may require more detailed investigation and flexibility in the approach to construction.
Some of the more notorious of these include: (1) karstic conditions associated with limestone and other rocks susceptible
to solution, (2) rock with steeply dipping discontinuities,
(3) well-developed residual soil deposits grading into partially
weathered rock and then unweathered bedrock, (4) alternating hard and soft layers of rock, and (5) glacial till. Each of
these conditions presents its own unique set of construction
challenges and different approaches are required to address
them successfully. A question that often arises in some of
these environments is what is rock?, or perhaps more importantly, what is not rock? On some projects, certain geomaterials may be rock for pay purposes, but not for design.
If these issues can be addressed before construction and there
is good communication between owners and contractors, a
reasonable approach that results in a successful project can
usually be developed. When the difficulties are not anticipated but are encountered during construction, the likelihood
of claims and disputes is much higher. Drilling of a trial installation shaft (also referred to as a method or technique
shaft) before bid letting can identify many of the problems
that will be encountered during production drilling and
should be considered whenever there are major questions

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

91

about the subsurface conditions and what is required to construct rock sockets successfully.
Inspection and Quality Assurance

Inspection is the primary method for assuring quality in the


construction of drilled shafts. The philosophy and methods
of drilled shaft inspection are covered in Chapter 16 of the
FHWA Drilled Shaft Manual (ONeill and Reese 1999) and
are the subject of a video and a Drilled Shaft Inspectors
Manual (Baker 1988) available from the ADSC. A certification course for drilled shaft inspectors is offered by the NHI
of FHWA, and a Participants Manual was developed as part
of the course (Williams et al. 2002). Table 21 is a partial listing of inspection issues pertaining specifically to rock-socket
construction.
Special emphasis is required in making a strong connection between drilled shaft design and inspection. Practically,
this involves providing inspection personnel with the knowledge and tools required to verify that drilled shafts are constructed and tested in accordance with the design intent. The
starting point for inspection personnel is to have a thorough
understanding of (1) subsurface conditions, (2) the intent of
the design, and (3) how items 1 and 2 are related. The inspectors sources of information for subsurface conditions

include the geotechnical report, boring logs, and communication with the design engineer. For rock sockets, inspectors
should be trained to understand the information presented in
boring logs pertaining to rock. This includes being familiar
with the site and geomaterial characterization methods described in chapter two. Inspectors require basic training in
rock identification, testing, and classification, and should be
familiar with rock coring procedures, the meaning of RQD,
compressive strength of intact rock, and terminology for
describing characteristics of discontinuities, degree of weathering, etc. Inspectors should be aware of design issues such
as whether the shaft is designed for side resistance, base resistance, or lateral resistance, and in which rock layers the
various components of resistance are derived.
Before construction, inspectors should know how the contractor plans to construct the shafts. This requires knowledge
of the tools and methods used for construction in rock. A
valuable aid is the Drilled Shaft Installation Plan, a document
describing in detail the contractors tools and methods of
construction. ONeill and Reese (1999) describe the minimum requirements of an installation plan and recommend
that it be a required submittal by the contractor.
A fundamental design issue is the degree to which the
rock mass over the depth of the socket coincides with the
conditions assumed for design. Therefore, some type of

TABLE 21
INSPECTION ITEMS FOR ROCK SOCKETS
Inspection Responsibility
Knowledge of site
conditions

Primary Items to Be Addressed


Rock types, depths, thicknesses, engineering
properties (strength, RQD); groundwater
conditions

Required Skills or Tools


Competency in rock identification and
classification; ability to read and interpret
core logs

Knowledge of design
issues

Rock units providing side, base, and lateral


resistances
Design parameters: shaft locations, socket
depths and diameters, reinforcement details

Basic understanding of design philosophy


for drilled shafts under axial and lateral
loading
Familiarity with standard specifications,
plans, special provisions, shop drawings,
and contractor submittals

Knowledge of contractors
plan for socket
construction

Rock excavation tools (augers, coring, hammers,


other) and methods (e.g., casing, slurry)
Classification of rock for pay purposes

Review of Drilled Shaft Installation Plan

Observations and record


keeping during socket
excavation

Identification and logging of excavated rock


Tools used by contractor for each
geomaterial (tool description, diameter, rate
of excavation)
Occurrence of obstructions, removal method
Depth to top of rock
Sidewall conditions (roughness, smearing)
Roughening or grooving of sidewalls
Use and handling of slurry and casing
Inspection methods and devices (e.g., SID)
Coring at the base
Cleanout specs., verification method

Competency in field identification of


geomaterials;
Appropriate forms*, including:
Rock/Soil Excavation Log
Rock Core Log
Inspection Log
Construction and Pay Summary

Sampling and testing

Sampling of rock for lab tests;


Proper sampling/testing equipment and
Field tests on rock; e.g., point load, hardness;
knowledge of procedures
NDT/NDE
*See Williams et al. (2002) for descriptions of inspection forms.
Notes: RQD = rock quality designation; SID = shaft inspection device; NDT = nondestructive testing; NDE =
nondestructive evaluation.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

92

downhole inspection is needed. Responses to Question 11


of the survey reveal a wide variety of methods used for this
purpose. Nine states reported that coring is required into rock
below the bottom of the shaft after the excavation to base
elevation is complete. Typical required depths range from
1.5 m to 10 m, three diameters, etc., although one state requires coring 15 m below the bottom of the shaft. Coring below the base during construction allows a determination to
be made of the adequacy of rock below the base to (1) provide the base resistance assumed in the design; (2) ensure
that the base is bearing on bedrock and not an isolated boulder (floater); and (3) detect the presence of seams, voids,
or other features that would require changes in the base
elevation or other remedial actions.
Five states reported using a probing tool to inspect core
holes at the bottom of the completed excavation (Figure 77).
This method, which in most cases requires downhole entry
by the inspector, is most useful for detecting seams of soft
material in discontinuities. It is most applicable in limestone
and dolomite where the bedrock surface is highly weathered,
irregular, and filled with slots and seams of clayey soil.
Proper safety measures are paramount for downhole entry.
Five states reported using fiber optic cameras for inspection
of core holes, which is safer and provides visual evidence of
seams, cavities, and fractures, but does not provide the feel
of probing that may be useful in karstic formations. Four of
the states reporting use of probe rods are in the Southeast
where karstic conditions are most common.

Most states include specifications for conditions at the


bottom of the hole that must be satisfied before pouring
concrete. Some distinguish between shafts designed for base
resistance and those designed under the assumption of zero
base resistance. A very typical specification (five states)
is minimum 50% of the base area to have less than 12 mm
(0.5 in.) and maximum depth not to exceed 38 mm (1.5 in).
Some states allow up to 300 mm (6 in.) of loose material
when base resistance is neglected.
When sockets are poured under dry conditions, common
inspection methods to verify bottom conditions are either
visual inspection or downhole cameras. For wet pours (under
slurry or water) the most common method is to lower a
weighted tape (e.g., a piece of rebar on the end of a tape measure) to the bottom of the hole and feel the bottom conditions by bobbing the weight against the bottom. Although
somewhat subjective, an experienced inspector can differentiate between clean water or slurry and contaminated conditions. Downhole cameras are available that permit viewing
of conditions under water or slurry. A device used by the
Florida DOT referred to as a shaft inspection device or SID
has been used successfully in slurry shafts (Crapps 1986).
The device, shown in Figure 78, has a color television cam-

Base of
Socket

Probe
Rod

FIGURE 77 Rock probing tool (after Brown 1990).

FIGURE 78 Shaft inspection device or SID.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

93

era encased in a watertight bell and equipped with a light


source and a water jet for clearing sediment to provide clear
pictures of the shaft sides and base. The SID was developed
in Australia specifically for inspection of rock sockets under
bentonite slurry. North Carolina also reported using a SID,
and several other states use downhole cameras to inspect
sockets under water or slurry.
The survey shows that some states neglect socket-base resistance altogether if concrete is placed under slurry or water
(Question 14). The rationale is that base conditions cannot be
verified with sufficient reliability to be sure that a poor base,
or soft bottom, condition is avoided. This refers to a layer
of disturbed soil, slurry, or contaminated concrete at the base,
which may allow excessively large downward movement before the resistance of the underlying rock can be mobilized.
These concerns may be justified under some conditions.
However, as described in chapter three (see Table 16 and
Figure 23), there are good reasons to account for base resistance even for shafts constructed under wet-hole conditions.
Construction and inspection practices that can be taken to
avoid poor base conditions include appropriate specifications
and quality control on properties of slurry at the bottom of the
hole prior to concrete placement, cleanout of slurry contaminated with cuttings or suspended particles before concrete
placement, use of a weighted tape to feel the bottom of the
hole as an inspection tool, downhole viewing devices for inspection of bottom conditions (e.g., SID), and proper use of
a pig or other device in the tremie pipe to prevent mixing
of concrete and slurry. Post-grouting of the shaft base is a
measure that could be incorporated into design and construction to provide quality base conditions in drilled shafts.
It is instructive to observe that most states that have
incorporated field load testing of rock sockets into their foundation programs, using a method that allows measurement
of base load-displacement, now include both side and base
resistances in their design calculations. This is based on load
test results that show, when proper quality control is applied,
that base resistance is a significant component of shaft resistance at service loads.

that may be of a smaller diameter than the shaft above the


socket.
EXAMPLES OF DIFFICULT GEOLOGIC
CONDITIONS

Some of the most difficult conditions for drilled shaft construction and inspection are karstic limestone and residual
profiles that grade from soil to weathered rock to intact rock.
Experiences and approaches to these conditions identified by
the literature review are summarized here.
Shafts in Limestone

Use of drilled shafts in karstic terrain is considered by Knott et


al. (1993), Sowers (1994), and others. Brown (1990) describes
design and construction challenges of using drilled shafts in hard
pinnacled limestones and dolomites encountered in the Valley
and Ridge and Cumberland Plateau physiographic provinces.
Subsurface conditions are highly irregular owing to extensive
weathering. Although intact rock strengths may be high (up to
70 MPa or 10,000 psi), numerous seams, slots, and cavities
are typically filled with residual clayey soils (see Figure 79).
Boulders and chert nodules are often embedded in the soils.
Drilling through soil is often performed in the dry soil and then
a casing set when rock is encountered. Drilling in the rock is
difficult and can involve a combination of rock augers, drill
and shoot methods, and core barrels. Sudden groundwater inflow is common upon encountering soil seams and slots.
In this environment of extreme variability the actual soil
and rock conditions for a specific drilled shaft cannot be
determined with any degree of accuracy before construction.
Design, construction, and inspection have to be flexible
enough to adjust to conditions actually encountered. For
example, where shafts can be shown to bear at least partially
on sound rock, base resistance is assumed, but highly

Nondestructive Testing and Evaluation

Field tests to evaluate the integrity of as-built drilled shafts are


now used widely in the industry as part of overall quality
assurance. Nondestructive methods for testing (NDT) and
evaluation (NDE) are covered in Chapter 17 of the Drilled
Shaft Manual (ONeill and Reese 1999) and in several other
publications. The survey for this study included a question
asking respondents to identify any issues pertaining to NDT
and NDE that are unique or important specifically for rocksocketed drill shafts. No issues were identified, other than the
need to consider locating NDT access tubes in the reinforcing
cage so that the entire assembly is able to fit into a socket

FIGURE 79 Features of karstic terrain (Knott et al. 1993).

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

94

(a)

(b)

(c)

FIGURE 80 Commonly encountered conditions for shafts in pinnacled limestone (Brown 1990).

conservative values are used to account for the presence of


seams at the base. This case is illustrated in Figure 80a, in
which a probe rod placed down one or more probe holes
drilled into the base can be used to determine the extent and
nature of the seam. One criterion for acceptance is rock coverage of 75% or more of the base area and vertical seams.
Figure 80b shows a nonvertical seam, which should be
detectable by one of the probe holes and might necessitate
additional drilling to preclude shear failure along the seam.
Alternatively, the seam could be excavated and grouted. This
technique would not be recommended if seepage is expected
into the excavated seam. Where shafts are bearing on a section of rock bounded by vertical seams or slots and the possibility of fracturing exists, rock anchors are sometimes used
to transfer load across potential fracture planes, as illustrated
in Figure 80c. Rock anchors or micropiles are also used to
transfer load across horizontal seams filled with soft soil and
detected by probing beneath the base.
To provide the flexibility needed for design, inspection,
and construction, creative contracting approaches are also
needed. Brown (1990) reported that contracting such work
on a unit cost basis provided the flexibility needed to deal
with the unknown quantities of soil versus rock drilling, concrete overpours, rock anchoring, drilling of probe holes, etc.
The engineer estimates the unit quantities, but actual payment is based on unit costs of material quantities actually
used. This requires careful inspection and record keeping.

Drilled Shafts in the Piedmont

The Piedmont Physiographic Province of the eastern United


States, extending from Alabama to New Jersey, is characterized

by decomposed metamorphic rocks and a weathering profile


characterized by unpredictable variability in the thickness and
quality of the weathered materials. Drilled shafts are used extensively for major structures in this region, primarily because
it has been recognized that large axial loads can be supported if
a shaft is extended to either decomposed or intact rock.
Gardner (1987) identified three general weathering horizons in the Piedmont: (1) residual soil, representing advanced
chemical alteration of the parent rock; (2) highly altered and
leached soil-like material (saprolite) retaining some of the
structure of the parent rock; and (3) decomposed rock (locally
referred to as partially weathered rock), which is less altered
but can usually be abraded to sand- and silt-sized particles.
The underlying intact rock is typically fractured near its surface but increases in quality with depth. The thickness and
characteristics of each zone vary considerably throughout the
region and may vary over short horizontal distances, and
boundaries between the horizons may not be distinct. Figure
81 shows a typical profile based on borings at one site. Factors
that make drilled shafts challenging to design and build in the
Piedmont are:

Highly variable subsurface profiles,


Presence of cobbles and boulders,
Steeply dipping bedrock surfaces, and
Difficulties in distinguishing between soil, partially
weathered rock, and intact rock for pay purposes.

The first of these makes it difficult to determine ahead of


time what the final base elevation will be for shafts required
to reach intact rock. At least one boring at each drilled shaft
location can help to address this issue. Figure 82 illustrates
two conditions that can cause refusal before the shaft

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

95
HORIZONTAL SCALE (FT)
0
930

100

300

B1
10
8
6
12

920
910

500

400

B3

B2

10

31

ZONE I

17

ZONE II

12
16

NX-18%
RQD-0

890

24

C.T.

870

NX-15%
RQD-0

CORING TERMINATED

11

-10

PENETRATION RESISTANCE

18

NX-18%

CORE RECOVERY

16

RQD-82

ROCK QUALITY DESIGNATION

19
53

ZONE III

50/3"
50/4"

NX-90%
RQD-79
C.T.

860

11

C.T.

10

50/5"
NX-95%
RQD-82

PARTIAL LEGEND

10

18
55
50/3"

900

880

200

ZONE IV

ZONE I

FILL

ZONE II

RESIDUAL SOIL

ZONE III

PARTIALLY WEATHERED ROCK

ZONE IV

ROCK

50/5"
NX-87%
RQD-10

GROUNDWATER, TIME OF BORING

24-HR GROUNDWATER

NX-95%
RQD-51

850

C.T.

ELEVATION (FT)

FIGURE 81 Typical Piedmont subsurface profile (after Schwartz 1987).

is drilled to its design base elevation. When refusal is encountered on a boulder that is floating, questions may
arise concerning whether the boulder is an obstruction or
constitutes drilling in rock. Similarly, when sloping bedrock
is first encountered, the volume of material excavated to
reach base elevation may be disputed as to whether it is soil
or rock, and drilling into sloping rock can be difficult. One
approach is to install casing until one edge of the casing hits
rock, then drill a smaller diameter pilot hole into the rock
followed by drilling to the design diameter and advancement
of the casing.

These examples illustrate the challenges that can be encountered in the design and construction of rock-socketed
drilled shafts as a result of certain geologic conditions, as
well as approaches that others have found successful for
addressing such challenges. Every foundation site is unique
geologically, and successful design and construction approaches are those that are adapted to fit the ground conditions. Mother Nature is quite unforgiving to those who behave
otherwise.

SUMMARY

Gardner (1987) reviews design methods for axial loading of drilled shafts in Piedmont profiles, including recommendations for design side and base resistances in rock and
methods used to determine relative load transfer between
side and base. Harris and Mayne (1994) describe load tests
in Piedmont residual soils. ONeill et al. (1996) used the
tests of Harris and Mayne to develop the recommendations
for side resistance in cohesionless IGM from Standard Penetration Test results, as presented in chapter three. Both
Gardner (1987) and Schwartz (1987) outline measures that
can be taken to minimize construction delays and contract
disputes when building rock-socketed shafts in Piedmont
profiles. The principal requirements are: (1) thorough site
investigation, (2) design and construction provisions that
can accommodate the unpredictable variations in subsurface materials and final base elevations, and (3) construction specifications and contract documents that facilitate
field changes in construction methods and shaft lengths.
Successful construction also depends on highly qualified
inspectors and clear communication between design engineers, contractors, and inspectors.

Construction and issues related to constructability are integral parts of drilled shaft foundation engineering. A review
of rock drilling technologies is presented and shows that a
wide variety of equipment and tools is available to contractors for building drilled shafts in rock. The design, manufacturing, and implementation of rock drilling tools is a field
unto itself and it is important for foundation designers to be
knowledgeable about the availability and capability of tools
and drilling machines. Constructability issues are interrelated
with all of the steps shown in the flowchart of Figure 3, depicting the design and construction process for rock-socketed
shafts. Beginning with site characterization and continuing
through final inspection, constructability is taken into account in foundation selection, in design methods through the
effects of construction on side resistance, in critical design
decisions such as whether base resistance will be included,
in writing of specifications pertaining to use of slurry and
bottom cleanout, and in matching inspection tools and procedures to construction methods. The literature review identified many aspects of constructability pertaining to rock

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

96

sockets and these are summarized. Practices that can improve


constructability; for example, the use of SCC and installation
of method shafts are identified.
Field load testing of rock sockets has increased since the
advent of innovative load testing methods, especially the
O-cell and the STN. The basic mechanics of these and other
tests are described, followed by a review of current applications of each to testing of rock-socketed shafts. The survey
shows that many states are using the O-cell to verify, and also
to improve, design methods of rock sockets. A description of
the KDOT experience with O-cell testing in rock is presented
as an example. Load testing is also shown to be a factor in
increased use of rock-socketed drilled shafts by transportation agencies. Finally, load testing with the O-cell has been
a useful tool for identifying and evaluating poor versus good
construction practices. The report by Schmertmann et al.
(1998), referenced several times in this chapter, is a particularly useful source for that information.
Inspection and field quality control are recognized in the
drilled shaft industry as the critical link between design and
construction. Excellent sources of information on inspection
are available and these are identified. The NHI inspector
certification course is highly recommended for all inspection personnel. Some of the tools identified by the survey
and literature review that can be most effective for rocksocket inspection are the SID, coring of rock beneath the
socket-base, use of probing tools, and downhole fiber optic
cameras.

FIGURE 82 Typical drilling in Piedmont soils and rock


(Schwartz 1987).

Two geologic environments in which rock-socket construction poses special challenges, karstic limestone and
Piedmont residual profiles, are presented to illustrate some
of the practices that lead to successful projects. Matching of
design and construction strategies to ground conditions is the
essence of constructability.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

97

CHAPTER SIX

CONCLUSIONS

This synthesis identified technologies and practices available to


transportation agencies for utilization of rock-socketed drilled
shafts as reliable and cost-effective structure foundations. All
thirty-two of the state transportation agencies responding to the
survey are currently using rock-socketed shafts, some quite extensively (more than 20 projects per year). The single Canadian
agency responding to the survey has not used drilled shafts
extensively to date. Use of rock-socketed drilled shafts for
transportation structures has increased significantly over the
past 25 years and technologies applied to design, construction,
and testing have advanced considerably.
The design process for structural foundations by state
transportation agencies is outlined in chapter one. Responsibilities typically are separated into geotechnical and structural categories. Site characterization, geomaterial property
evaluation, and design issues related to geotechnical capacity or load-deformation analysis are normally addressed as
geotechnical issues, whereas structural modeling and reinforced-concrete design are normally carried out by structural
engineers. Design for lateral loading requires significant input
and analysis by both geotechnical and structural personnel.
The overall process of design and construction (i.e., engineering) is shown to consist of highly interrelated factors,
requiring an integrated approach to drilled shaft foundations.
Figure 3 in chapter one illustrates the process in the form of
a flowchart. Adequate site characterization is needed to
obtain the basic information required for both geotechnical
analysis and construction. Constructability issues are best
addressed during the design process, when decisions such as
whether to include side resistance, base resistance, or both
must be made on the basis of anticipated subsurface conditions, construction methods, load testing, inspection methods, and experience.

SITE AND GEOMATERIAL CHARACTERIZATION

The most valuable and reliable information for rock-socket


design is obtained by drilling and taking core samples of the
rock at the location of each structural foundation. Careful
logging of rock core, photographic records, and proper handling of core to obtain samples for laboratory testing provide
the basic information that will be used for rock mass classification, evaluation of engineering properties of intact rock

and rock mass, and baseline information needed to assess


constructability. Drilling also provides the means to conduct
in situ tests. Every transportation agency that responded to
the survey currently relies on rock coring as the primary source
of design and construction information for rock-socketed
shafts.
Geophysical methods can provide additional valuable information when applied appropriately by competent users.
NCHRP Synthesis 357: Use of Geophysics for Transportation Projects (Sirles) identifies the major geophysical methods
that are applicable to geotechnical investigations and found
that overall use of geophysics by transportation agencies is expanding. Seismic refraction for establishing depth to bedrock
is the most common use of geophysics for drilled shafts in
rock. However, of 33 responding agencies, only 8 (24%) reported using geophysics, including 7 that use seismic refraction and 1 that uses electrical resistivity. These data suggest
that geophysical methods are not used widely for investigations related specifically to foundations in rock. Survey
results from the Sirles study show that agency experience is
mixed, with both successful and unsuccessful cases being
cited. Factors associated with successful cases (for depth to
bedrock) are: sufficient number of borings to validate and
correlate the seismic results, interpretation by qualified geophysicists, and clear understanding of the capabilities and
limitations of the technology.
Geophysical methods that show potential for rock site investigations include electrical resistivity tomographic profiling and borehole televiewers. Multi-array resistivity methods
have shown the ability to provide accurate images of subsurface profiles in karstic terrains when used in conjunction
with borings. Borehole televiewers, both acoustic and optical, may have limited applicability to rock foundations. They
are primarily useful for providing detailed information on
structural discontinuities. For large or critical rock-socket
projects, where the orientation and condition of discontinuities in situ is a critical concern, these devices can supplement
information obtained from more conventional core logging.
Other potentially useful methods are downhole seismic and
crosshole seismic. A case described by LaFronz et al. in
2004, Geologic Characterization of Bridge Foundations,
Colorado River Bridge, Hoover Dam Bypass Project, showed
good correlation between rock mass modulus from downhole
seismic testing and rock mass modulus from correlations to

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

98

the Geological Strength Index (GSI). This approach warrants


further consideration.
The laboratory test most widely applied to foundation
design is uniaxial compression of intact rock. Properties
obtained are uniaxial compressive strength (qu) and elastic
modulus of intact rock (ER). Poissons ratio may also be
determined. Uniaxial compressive strength is used directly
in the most widely applied design methods for evaluating
unit side resistance, unit base resistance, and limiting pressure under lateral loading. Modulus of intact rock is not
used directly, but rather with other rock mass characteristics to evaluate rock mass deformation modulus (EM).
Other laboratory tests applicable to rock-socket design include the splitting tensile strength (used for side resistance
in limestone) and the point load strength (an index of compressive strength). Direct shear testing is used to assess
shear strength of rock mass discontinuities and can be used
to test shear strength of rock/concrete interfaces. Slake
durability is used to assess potential for rapid degradation
and smearing of weak rocks during construction of rock
sockets.
Rock mass classification systems have useful applications
in foundation design. The Rock Mass Rating (RMR) as given
by Bieniawski in Engineering Rock Mass Classifications
(1989) incorporates the most important rock mass characteristics (including rock quality designation) that control the
strength and deformability of a rock mass. The RMR is useful as an overall indicator of rock quality and suitability as a
founding material, and is the basis for correlations to rock
mass strength and modulus. Approximately one-half of the
states responding to the survey reported using RMR in connection with rock-socket projects. The GSI introduced by
Hoek et al. in Support of Underground Excavations in Hard
Rock (1995) can be evaluated on the basis of RMR and is
also correlated directly with rock mass strength, through the
HoekBrown strength criterion, and rock mass modulus of
deformation. GSI is now being used in geomechanical
models for bearing capacity in rock and for evaluation of
limiting lateral pressure for shafts in rock under lateral
loading.
In situ testing in rock is used primarily to obtain rock mass
modulus (EM). Pressuremeter (PMT) and borehole jack are
the methods being used. Modulus values obtained by PMT
are affected by the scale of the test relative to the scale of
rock mass features (discontinuity spacing and orientation)
and may or may not be representative for the purpose of
foundation analysis. The principal use of rock mass modulus
is in analyzing axial and lateral load deformation response of
rock sockets. There are several published p-y curve criteria
for laterally loaded shafts that incorporate modulus as determined by PMT. The issue of whether the modulus values
from PMT are the most appropriate requires further research.
Table 15 in chapter two summarizes the rock mass properties
required for design of rock-socketed shafts.

DESIGN FOR AXIAL LOAD

Methods for predicting the behavior of rock sockets under


axial loading have developed considerably since the 1970s.
The literature review showed that axial load transfer is reasonably well understood in terms of its basic mechanisms.
Effects of interface roughness, socket-length-to-diameter
ratio, modulus ratio, and other variables have been studied
analytically and experimentally, providing a broad understanding of the underlying concepts. Although design methods do not incorporate all of the governing parameters
explicitly, understanding the underlying mechanics is useful
in many ways, including to provide a framework for understanding the limitations of empirical design methods.
Specific methods for predicting side and base capacities
must be in a form that matches the level of knowledge of the
ground conditions and that is based on commonly measured
rock and intermediate geomaterials (IGM) properties. Chapter three of this synthesis provides a summary and review of
available methods and it is shown that conservative, reliable,
first-order estimates can be made for design values of side
resistance on the basis of uniaxial strength of intact rock.
Geomaterial-specific methods are presented for Florida limestone, residual Piedmont soils (cohesionless IGMs), and
weak argillaceous rocks (cohesive IGMs). A method based
on direct correlation to Texas Cone Penetration Test results
illustrates how some agencies use in-house methods.
For base capacity, a variety of methods have been proposed in the literature. Because several modes of failure
are possible depending on structural characteristics of the
rock mass, no single equation is applicable to all conditions.
Furthermore, few studies have been conducted comparing
proposed bearing capacity models with measured base capacities on socketed shafts loaded to failure. A 1998 study by
Zhang and Einstein provided a first-order approximation of
unit base resistance from uniaxial strength of intact rock,
based on a limited database of field load tests. For intact rock,
a conservative upper-bound unit base resistance can be taken
as 2.5 times the uniaxial compressive strength. Two methods
given in the Canadian Foundation Engineering Manual
and incorporated into current AASHTO specifications are
recommended for horizontally jointed sedimentary rocks.
For highly fractured rock masses, a lower-bound estimate
of ultimate bearing capacity can be made in terms of RMR
or GSI.
Analytical methods for predicting axial load-displacement
of rock sockets are needed to design shafts to limit settlement
and to determine the percentage of load carried by base resistance under service load conditions. Methods based on elastic and elastoplastic finite-element modeling are available in
the form of charts. Although these methods are useful, they
are cumbersome. Simple closed-form solutions that are
implemented easily on a spreadsheet are presented. Both
approaches require knowledge of the rock mass modulus.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

99
DESIGN FOR LATERAL LOAD

Methods for analysis of rock sockets under lateral loading are


readily available to foundation designers, but currently are
subject to uncertainties regarding their reliability. The survey
shows that all states currently use the p-y method of analysis.
Criteria for p-y curves in rock have been proposed and these
are the most widely used at the current time. However, the
proposed criteria were described as interim when they first
appeared, because of the insufficient field load test data available for validation. Research aimed at improving p-y curve
criteria for rock has been described. The proposed methods
also require additional verification by comparisons with field
load testing. A major feature of p-y methods of analysis is that
they provide structural analysis of the reinforced-concrete
shaft that incorporates the nonlinear momentEI relationship.
This feature provides a useful interface between geotechnical
and structural design.
Analysis methods based on elastic continuum theory have
been developed for lateral loading. The Carter and Kulhawy
method (1992) requires a minimal number of parameters and
is easy to implement by hand or spreadsheet, but is applicable only over the range of elastic response. The Zhang et al.
method (2000) provides the complete nonlinear response, but
requires more input parameters and relies on a finite-difference
computer solution. These methods may be useful in the Preliminary Foundation Design phase (Figure 2, chapter one),
for making first-order assessments of trial designs to satisfy
service limit state criteria for lateral displacements. They are
most applicable when the ground profile can be idealized as
consisting of one or two homogeneous layers; for example,
soil over rock.

LOAD TESTING

A positive development for drilled shaft design has been the


introduction of several innovative field load testing methods.
The Osterberg Cell (O-cell) and Statnamic (STN) tests can
be conducted in less time, at lower cost, and with less equipment than conventional axial load testing methods. This has
given transportation agencies the option of incorporating
load testing into the design process on individual projects
and developing databases of shaft performance in specific
geologic environments. The experience of the Kansas Department of Transportation is described as a model example
for incorporating O-cell testing into a comprehensive program that has resulted in more efficient use and design of
rock-socketed shafts. Many of the states surveyed have taken
advantage of O-cell and STN testing and this has resulted in
a significant increase in load test data. It is suggested that a
database of load test results be developed, analyzed, maintained, and made available to the wider research community.
The survey shows that states using the O-cell for axial
load testing are less likely to neglect base resistance for

design of rock-socketed shafts. O-cell testing almost always


demonstrates that base resistance provides a significant portion of total axial resistance under service load conditions.
Data from Crapps and Schmertmann in Figure 23, chapter
three, show direct evidence of significant base load transfer
when appropriate construction and inspection methods are
applied to base conditions. Furthermore, O-cell and STN
testing often result in higher values of allowable side resistance than would be calculated using the recommended
prediction equations, which are intended to be conservative.
Lateral load testing of rock sockets can be conducted
using O-cell and STN methods. The STN method may be
particularly applicable for design of shafts subject to
dynamic lateral loading, such as impact and seismic. Lateral
O-cell testing has been demonstrated successfully, although
research is suggested to develop procedures to relate lateral
O-cell test results to p-y curve criteria and to parameters used
in other analytical methods. Conventional static lateral load
testing is still the most common method and is a proven
approach to verifying performance and studying load transfer behavior. Lateral load testing on instrumented shafts is
the only reliable method for validating p-y curves for design.

CONSTRUCTABILITY AND INSPECTION

Issues of constructability and inspection are related directly


to rock-socket design and performance. Load testing, especially with O-cell methods, has helped to identify the effects
of various construction methods on rock-socket performance. For example, the perception that construction of rock
sockets is best facilitated by using full-depth casing and taking measures to permit a dry pour has been shown to have
detrimental effects on side and base resistances. Use of water
or slurry, when subjected to appropriate quality control, provides better performance by eliminating inward seepage,
trapping of cuttings behind casing, and potential for smearing as the casing is removed.
Tools available for incorporating constructability into
rock-socket design through specifications, plans, and inspection procedures are identified in several publications, including the FHWA Drilled Shaft Manual and the Participants
Manual for the National Highway Institute Inspectors Certification Course. Several state agencies have developed
model drilled shaft specifications that incorporate proven
constructability practices (see for example, Washington State
DOT Geotechnical Design Manual 2005). Recent developments in concrete mix design, such as self-consolidating
concrete, are expected to provide improved constructability.
Inspection tools such as the shaft inspection device used
by Florida and North Carolina have direct implications for
design. By providing a means to verify base conditions under water or slurry construction, designers are better prepared
to include base resistance in socket design. Construction of
technique or method shafts and contractor constructability

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

100

reviews before publication of the final design and bidding


phases are additional tools for incorporating constructability.
RESEARCH NEEDED TO ADVANCE
STATE OF PRACTICE

Information gathered for this study suggests that development of improved practices for design and construction
of rock-socketed drilled shafts might be achieved through
the following research or wider dissemination of existing
information.
Site Characterization

Studies are needed to better define the best methods for


determination of rock mass deformation modulus specifically for use in rock-socket design. In situ methods,
including borehole jack and PMT, may yield different
results and both could be compared with the most up-todate correlations with RMR and GSI.
A survey of contractors could be conducted to identify
the rock mass information most useful for evaluating
construction in rock; avoiding overemphasis on weakest materials.
Application of geophysical methods to rock-socket
design requires further research and development.
Guidelines are needed for matching appropriate methods to site conditions. Case histories of successes and
limitations could be published and distributed.
Research is needed relating rock drillability to rock mass
characteristics; correlations to RMR or GSI warrant
investigation.
Relationships between the reliability of rock and IGM
engineering properties and resistance factors used in
load and resistance factor design could be investigated
and quantified sufficiently to support the resistance values recommended in AASHTO specifications; this
topic could be the subject of ongoing research.

proposed by several researchers. These design methods are


limited because roughness is not a commonly measured parameter in the field. Construction techniques are constantly
under development and innovative methods that can lead to
improved quality should be encouraged and, where appropriate, developed further through research.
Consider developing a manual or design circular focused specifically on drilled shafts in rock.
Research is needed for axial load tests on instrumented
shafts for the purpose of evaluating prediction equations for base resistance in rock; O-cell and STN tests
are ideal for this purpose.
Identify efficient and inexpensive field roughness measurement methods that can be incorporated into design
equations; correlate roughness parameters to rock type,
drilling tools, groundwater conditions, etc.
Investigate the potential of base grouting as a quality
assurance tool for rock-socketed shafts.
Design for Lateral Loading

Methods developed for analysis of deep foundations in soil,


especially the p-y curve method, are the methods of choice
for laterally loaded rock sockets. The principal limitation lies
in the lack of proven p-y curve criteria for rock and IGM.
This problem could be addressed by first conducting a comprehensive analysis of all existing load test results to evaluate
proposed models, followed by research involving additional
field load testing against which p-y curve criteria can be evaluated and calibrated.
Conduct research to collect and analyze all existing
lateral load test results, with the goal of establishing
uniform criteria for p-y curve development.
Transportation agencies could undertake research involving lateral load tests on properly instrumented
rock-socketed shafts, designed specifically for testing
and calibration of p-y criteria for rock and IGM.

Design for Axial Loading

Structural Design

Sufficient analytical tools exist for the reliable design of


sockets under axial loading. However, much of this information is widely scattered in the literature. The FHWA Drilled
Shaft Manual and the 2006 Interim AASHTO LRFD Bridge
Design Specifications include some available methods, but
are not concise and clear in the presentation, and include
some out-of-date methods. Numerous equations are presented
in the literature for estimating base resistance of drilled shafts
in rock. Surprisingly, very little data are available by which
proposed methods can be evaluated. Studies are needed involving field axial load testing in which rock mass properties
are well-documented and design equations for base resistance can be evaluated. Equations for incorporating roughness as a design parameter for unit side resistance have been

Structural issues of concern to foundation designers, as identified by the survey, included uncertainty regarding apparently high shearing forces in shafts analyzed using p-y curve
analyses and questions pertaining to moment capacity of
short, rigid sockets. These issues can best be addressed by
rigorous analytical methods in conjunction with load tests on
carefully instrumented shafts in rock. A structural issue that
has yet to be investigated as it pertains to deep foundations is
the effect of confining stress on the strength and stiffness of
reinforced concrete. It may be that concrete strength could be
significantly increased under confining stresses typically encountered over the subsurface depths of many bridge foundations. More economical structural designs may be possible
if this issue is investigated and applied in practice. Permanent

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

101

steel casing contributes to the structural capacity of drilled


shafts. Design methods that account explicitly for the steel
casing are lacking in current design codes.
Consider fundamental research with the goal of quantifying the effects of geologic confining stress on reinforcedconcrete shear, moment, and compression behavior.
Incorporate the results into structural design of drilled
shafts.
Conduct research and development of methods that account for permanent steel casing in the structural design
of drilled shafts.

Management of Load Test Data

Large amounts of data from load tests on rock-socketed


shafts, conducted for the purpose of research or for specific
transportation projects, have been acquired, especially since
the development of new testing methods. These data can
be used most effectively if they are made available from a
single source and organized in a systematic manner.
Investigate placing those into a national database of
load test results for rock-socketed drilled shafts, for use
by transportation agencies and researchers.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

102

EQUATION SYMBOLS

ENGLISH LETTERSUPPERCASE

Ab = cross-sectional area of the foundation base.


Ac = cross-sectional area of concrete inside of spiral
steel.
Ag = gross cross-sectional area of concrete shaft.
As = surface area of the side of the foundation.
Ast = total cross-sectional area of longitudinal reinforcement.
Av = area of concrete in the cross section that is effective in resisting shear.
Avs = cross-sectional area of shear reinforcement.
B = foundation diameter.
Bls = diameter of a circle passing through the center of
the longitudinal reinforcement.
C = correlation factor relating point load strength to
uniaxial compressive strength of rock;
correlation coefficient relating uniaxial compressive strength to ultimate unit side resistance.
D = foundation depth; distance between point loads
in point load test; rock-socket diameter.
D s = depth of embedment in rock; thickness of soil
layer overlying rock.
E = elastic modulus; modulus of deformation.
Eb = elastic modulus of rock mass below the shaft
base.
Ec = elastic modulus of the concrete shaft.
Ed = small-strain dynamic modulus.
Ee = effective elastic modulus of concrete shaft.
Eir = initial elastic modulus of rock mass.
EM = rock mass modulus of deformation.
Ep = elastic modulus of the pile (shaft) material.
Er = elastic modulus of intact rock; modulus of rock
mass above the base.
ER = elastic modulus of intact rock.
Es = elastic modulus of the shaft.
G* = equivalent shear modulus of rock mass.
Gr = shear modulus of the elastic rock mass.
GSI = Geological Strength Index
H = horizontal load acting on a drilled shaft.
ID = slake durability index.
IGM = intermediate geomaterial.
Is = uncorrected point load strength index; moment of
inertia of reinforced-concrete shaft.
Is(50) = point load strength corrected to a diameter of
50 mm.
I = dimensionless influence factor for elastic
deformation.
J = coefficient used to evaluate ultimate lateral resistance in soil; bearing capacity correction
factor that depends on the ratio of horizontal discontinuity spacing to socket diameter.
Ja = joint alteration number.

Jn
Jr
Jw
K
Kb
Kir
Ko
Kp
L
LM

=
=
=
=
=
=
=
=
=
=

LV =
LM =
LV =
Ls =
Lt =
M =
N =
N60 =
Nc
Ncr
Nq
N
N
OCR
P

=
=
=
=
=
=
=

Pr
Pu
Pz
Q
Q'
Qb

=
=
=
=
=
=

Qc =
Qi =
QOC =
Qs =
Qt =
REC
RF
RMR
RQD
SRF
S

=
=
=
=
=
=

joint set number.


joint roughness number.
joint water reduction factor.
normal stiffness of rock-concrete interface.
socket depth factor.
initial slope of p-y curve.
in situ coefficient of lateral earth pressure.
coefficient of passive earth pressure.
socket length.
length of equivalent fixed-end column considering lateral deflection owing to moment.
length of equivalent fixed-end column considering lateral deflection owing to shear.
length of equivalent fixed-end column considering rotation owing to moment.
length of equivalent fixed-end column considering rotation owing to shear.
nominal socket length.
total travel distance along socket wall profile for
roughness determination.
bending moment.
standard penetration test value.
corrected N for field procedures corresponding to
60% hammer efficiency.
bearing capacity factor.
bearing capacity factor.
bearing capacity factor.
bearing capacity factor.
bearing capacity factor.
overconsolidation ratio.
axial load acting on a drilled shaft; load at rupture
in a point load test.
factored axial resistance.
factored axial load.
axial load on a deep foundation.
Tunneling Quality Index.
modified Tunneling Quality Index.
load transmitted to the base of a rock socket; ultimate resistance at the socket base.
compressive force applied to the top of a drilled
shaft.
intercept on the vertical axis (wc = 0) of axial
load-displacement curve.
Osterberg cell test load.
total side resistance (force).
total compressive load applied to the top of the
shaft.
average percent recovery of rock core.
roughness factor.
rock mass rating.
rock quality designation.
stress reduction factor.
joint spacing.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

103

S1 = slope of the initial portion of axial load-displacement curve.


S2 = slope of the full-slip portion of axial load-displacement curve.
Vc = nominal shear resistance provided by concrete.
Vn = nominal shear resistance of reinforced concrete.
Vp = compressional wave velocity.
Vr = factored shear resistance of reinforced concrete.
Vs = shear wave velocity; nominal shear resistance
provided by transverse steel.
ENGLISH LETTERSLOWERCASE

a = empirical constant in HoekBrown strength criterion for rock mass.


b = empirical factor to account for effect of roughness on side resistance.
c = rock mass cohesion; soil cohesion.
cpeak = peak interface adhesion.
cresidual = residual interface adhesion.
cu = undrained shear strength.
d = effective shear depth of reinforced concrete.
ds = socket diameter.
f = shear wave frequency (hertz).
fc' = compressive strength of concrete at 28 days.
fdes = design unit side resistance.
fsu = ultimate unit side resistance.
fy = yield strength of reinforcing steel.
kir = dimensionless constant used in p-y curve
criterion.
krm = constant used to establish the overall stiffness of
a p-y curve.
mb = empirical constant in HoekBrown strength criterion for rock mass.
n = ratio of rock mass modulus to uniaxial compressive strength of intact rock.
p = lateral soil or rock reaction per unit length of
foundation.
p1 = limit pressure determined from a pressuremeter
test.
pa = atmospheric pressure.
pA = horizontal active earth pressure.
pL = limit normal stress.
po = at-rest total horizontal stress.
pur = rock mass ultimate lateral resistance.
pult = ultimate lateral resistance of soil or rock mass.
qu = uniaxial compressive strength of intact rock.
qult = ultimate bearing capacity.
qt = split tensile strength.
r = radius of drilled shaft.
rs = nominal socket radius.
s = empirical constant in HoekBrown strength criterion for rock mass; vertical spacing of the ties
or pitch of the spiral for shear reinforcement.
sc, s, sq = shape factors used in bearing capacity analysis.
sv = vertical spacing between discontinuities.
td = aperture (thickness) of discontinuities.

u = lateral displacement at the groundline of socketed


shaft.
w = distributed load along the length of the shaft.
wc = axial displacement at the top of a socketed shaft.
xr = depth below rock surface.
y = lateral deflection of a deep foundation.
z = depth below rock mass surface.
zy = depth of yielding in soil and/or rock mass.
GREEK SYMBOLSUPPERCASE

r = dilation or increase in shaft radius.


r = mean roughness height.
rh = average height of asperities.
GREEK SYMBOLSLOWERCASE

= ratio of design to ultimate unit side resistance; ratio


of rock mass modulus to modulus of intact rock;
empirical adhesion factor relating unit side resistance to uniaxial strength of intact rock.
= total unit weight of rock or soil.
' = effective unit weight of rock or soil.
r = relative shear displacement between concrete and
rock.
M = lateral deflection owing to moment.
V = lateral deflection owing to shear.
= strain.
= Poissons ratio.
c = Poissons ratio of concrete.
b = Poissons ratio of rock mass beneath the base of
a rock socket.
r = Poissons ratio of rock mass.
c = construction method reduction factor.
= mass density.
s = ratio of spiral steel reinforcement volume to volume of concrete core.
'1, '3 = major and minor principal effective stresses.
'h = horizontal effective stress.
n = interface normal stress; fluid pressure exerted by
concrete at the time of placement.
p' = preconsolidation stress.
v = vertical total stress.
v' = vertical effective stress,
= shear strength.
max = shearing resistance at shaft-rock interface.
= shaft rotation.
M = rotation owing to moment.
V = rotation owing to shear.
= friction angle; resistance factor.
' = effective stress friction angle.
peak = peak interface friction angle.
rc = residual angle of interface friction between rock
and concrete.
residual = residual friction angle.
= angle of dilation.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

104

REFERENCES

Abu-Hejleh, N., M.W. ONeill, D. Hanneman, and W.J. Atwooll, Improvement of the Geotechnical Axial Design
Methodology for Colorados Drilled Shafts Socketed
in Weak Rocks, Report No. CDOT-DTD-R-2003-6,
Colorado Department of Transportation, Denver, 2003,
192 pp.
Acker, W.L., III, Basic Procedures for Soil Sampling and
Core Drilling, Acker Drill Company, Inc., Scranton, Pa.,
1974, 246 pp.
Allen, T.M., Development of Geotechnical Resistance Factors and Downdrag Load Factors for LRFD Foundation
Strength Limit State Design, Publication FHWA-NHI05-052, Federal Highway Administration, Washington,
D.C., Feb. 2005.
Annual Book of ASTM Standards, American Society for
Testing and Materials, West Conshohocken, Pa., 2000.
Amir, J.M., Piling in Rock, Balkema, Rotterdam, The
Netherlands, 1986.
Ashour, M., G. Norris, and P. Pilling, Lateral Loading of a
Pile in Layered Soil Using the Strain Wedge Model,
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 124, No. 4, 1998, pp. 303315.
Ashour, M. and G. Norris, Modeling Lateral Soil-Pile Response Based on Soil-Pile Interaction, Journal of Geotechnical and Geoenvironmental Engineering, Vol. 126,
No. 5, 2000, pp. 420428.
Ashour, M., G. Norris, S. Bowman, H. Beeston, P. Pilling,
and A. Shamabadi, Modeling Pile Lateral Response in
Weathered Rock, Proceedings, Symposium on Engineering Geology & Geotechnical Engineering, University of Nevada, Las Vegas, 2001, pp. 639649.
Barker, R.M., J.M. Duncan, K.B. Rojiani, P.S.K. Ooi, C.K.
Tan, and S.G. Kim, NCHRP Report 343: Manuals for the
Design of Bridge Foundations, Transportation Research
Board, National Research Council, Washington, D.C.,
1991, 308 pp.
Barton, N., Review of a New Shear Strength Criterion for
Rock Joints, Engineering Geology, Vol. 7, 1973, pp.
287332.
Barton, N., R. Lien, and J. Lunde, Engineering Classification of Rock Masses for the Design of Tunnel Support, Rock Mechanics, Vol. 6, No. 4, 1974, pp.
183236.
Basic Geotechnical Description of Rock Masses, International Journal Rock Mechanics, Mineral Science and Geomechanics Abstracts, Vol. 18, No. 1, 1981, pp. 85110.
Bedian, M.P., Value Engineering During Construction,
Geotechnical Special Publication No. 124: Geo-Support
2004, 2004, pp. 5269.
Bermingham, P., C.D. Ealy, and J.K. White, A Comparison
of Statnamic and Static Field Tests at Seven FHWA
Sites, Proceedings, Fifth International Deep Founda-

tions Institute Conference, Bruges, Belgium, 1994, pp.


616630.
Bieniawski, Z.T., Geomechanics Classification of Rock
Masses and Its Application in Tunnelling, Proceedings,
Symposium on Exploration for Rock Engineering,
Balkema, Rotterdam, The Netherlands, 1976, pp. 97106.
Bieniawski, Z.T., Determining Rock Mass Deformability
Experience from Case Histories, International Journal
of Rock Mechanics and Mineral Science, Vol. 15, 1978,
pp. 237247.
Bieniawski, Z.T., Rock Mass Design in Mining and Tunneling, Balkema, Rotterdam, The Netherlands, 1984, 272 pp.
Bieniawski, Z.T., Engineering Rock Mass Classifications,
Wiley, New York, 1989.
Bishnoi, B.W., Bearing Capacity of Jointed Rock, Ph.D.
thesis, Georgia Institute of Technology, Atlanta, Ga.,
1968.
Bloomquist, D. and F.C. Townsend, Development of In Situ
Equipment for Capacity Determinations of Deep Foundations in Florida Limestones, Report to Florida Department of Transportation, University of Florida, Gainesville,
1991.
Bridge Design Aids, Massachusetts Highway Department,
Boston, 1990.
Broms, B.B., Lateral Resistance of Piles in Cohesive Soils,
Journal of the Soil Mechanics and Foundations Division,
Vol. 90, No. SM2, 1964a, pp. 2763.
Broms, B.B., Lateral Resistance of Piles in Cohesionless
Soils, Journal of the Soil Mechanics and Foundations
Division, Vol. 90, No. SM3, 1964b, pp. 123156.
Brown, D.A., Construction and Design of Drilled Shafts in
Hard Pinnacled Limestones, Transportation Research
Record 1277, Transportation Research Board, National
Research Council, Washington, D.C., 1990, pp.
148152.
Brown, D.A., Evaluation of Static Capacity of Deep Foundations from Statnamic Testing, Geotechnical Testing
Journal, Vol. 17, No. 4, 1994, pp. 404414.
Brown, D.A., Load Testing for Drilled Shaft Foundations,
Lecture Notes, Faculty Workshop 2000, International
Association of Foundation Drilling, Dallas, Tex., 2000.
Brown, D.A., Zen and the Art of Drilled Shaft Construction:
The Pursuit of Quality, Geotechnical Special Publication No. 124: Geo-Support 2004, J.P. Turner and P.W.
Mayne, Eds., 2004, pp. 1933.
Brown, D.A., S.A. Hidden, and S. Zhang, Determination of
p-y Curves Using Inclinometer Data, Geotechnical Testing Journal, Vol. 17, No. 2, 1994, pp. 150158.
Building Code Requirements for Structural Concrete,
ACI318-94, American Concrete Institute, Farmington
Hills, Mich., 1985.
Bukovansky, M., Determination of Elastic Properties of
Rocks Using Various Onsite and Laboratory Methods,

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

105

Proceedings, 2nd Congress of the International Society of


Rock Mechanics, Belgrade, Yugoslavia, Vol. 1, 1970, pp.
329332.
Butterfield, R. and P.K. Banerjee, The Elastic Analysis of
Compressible Piles and Pile Groups, Geotechnique, Vol.
21, No. 1, 1971, pp. 4360.
Byles, R., AGRA Foundations and Seacore Combine to
Strengthen Major US Bridge, Foundation Drilling, Vol.
25, No. 7, International Association of Foundation
Drilling, Dallas, Tex., 2004, pp. 1013.
Caltrans, Sacramento, Calif., 2005 [Online]. Available:
http://www.dot.ca.gov/hq/esc/geotech/gg/atv_log.htm
[June 2005].
Canadian Foundation Engineering Manual, 2nd ed.,
Canadian Geotechnical Society, Ottawa, Canada, 1985,
456 pp.
Carter, J.P. and F.H. Kulhawy, Analysis and Design of
Drilled Shaft Foundations Socketed into Rock, Report
EL-5918, Electric Power Research Institute, Palo Alto,
Calif., 1988, 188 pp.
Carter, J.P. and F.H. Kulhawy, Analysis of Laterally
Loaded Shafts in Rock, Journal of Geotechnical Engineering, Vol. 118, No. 6, 1992, pp. 839855.
Cavusoglu, E., M.S. Nam, M.W. ONeill, and M. McClelland, Multi-Method Strength Characterization for Soft
Cretaceous Rocks in Texas, Geotechnical Special
Publication No. 124: Geo-Support 2004, J.P. Turner and
P.W. Mayne, Eds., 2004, pp. 199210.
Cokceoglu, C., H. Sonmez, and A. Kayabasi, Predicting the
Deformation Moduli of Rock Masses, International
Journal of Rock Mechanics and Mining Sciences, Vol. 40,
2003, pp. 701710.
Coon, R.F. and A.T. Merritt, Predicting In Situ Modulus of
Deformation Using Rock Quality, Special Technical
Publication No. 477, American Society for Testing and
Materials (ASTM), Philadelphia, Pa., 1969.
Crapps, D.K., Design, Construction, and Inspection of
Drilled Shafts in Limerock and Limestone, Proceedings,
35th Annual Geotechnical Conference, ASCE/AEG, University of Kansas, Lawrence, 1986, 38 pp.
Crapps, D.K., Proposed Improvements for Drilled Shaft
Design in Rock, Prepared for the Florida Department of
Transportation by Schmertmann & Crapps, Inc.,
Gainesville, Fla., 2001, 37 pp.
Crapps, D.K. and J.H. Schmertmann, Compression Top
Load Reaching Shaft BottomTheory Versus Tests,
Proceedings, International Deep Foundations Congress,
Orlando, Fla., Feb. 1416, 2002.
Davisson, M.T., Lateral Load Capacity of Piles, Highway
Research Record 333, Highway Research Board,
National Research Council, Washington, D.C., 1970, pp.
104112.
Deere, D.U., Technical Description of Rock Cores for
Engineering Purposes, Rock Mechanics and Engineering Geology, Vol. 1, No. 1, 1964, pp. 1722.
Deere, D.U. and D.W. Deere, Rock Quality Designation
(RQD) After Twenty Years, Contract Report GL-89-1,

U.S. Army Engineer Waterways Experiment Station,


Vicksburg, Miss., 1989.
DiMaggio, J.A., Developments in Deep Foundation Highway PracticeThe Last Quarter Century, Foundation
Drilling, Vol. 24, No. 2, International Association of
Foundation Drilling, Dallas, Tex., 2004, pp. 1630.
DiMillio, A.F., A Quarter Century of Geotechnical
Research, Report FHWA-RD-98-13X, Federal Highway
Administration, Washington, D.C., 1998, 160 pp.
Donald, I.B., S.W. Sloan, and H.K. Chiu, Theoretical
Analyses of Rock Socketed Piles, Proceedings, International Conference on Structural Foundations on Rock,
Vol. 1, Sydney, Australia, 1980, pp. 303316.
Dunscomb, M.H. and E. Rehwoldt, Two-Dimensional Profiling: Geophysical Weapon of Choice in Karst Terrain
for Engineering Applications, Hydrogeology and Engineering Geology of Sinkholes and KarstProceedings of
the Seventh Multidisciplinary Conference on Sinkholes
and the Engineering and Environmental Impacts of Karst,
Hershey, Pa., 1999, pp. 219224.
Eliassen, T., D. Richter, H. Crow, P. Ingraham, and T. Carter,
Use of Optical Televiewer and Acoustical Televiewer
Logging Methods in Lieu of Oriented Coring Methods,
Preprint, TRB Workshop on Geotechnical Methods
Revisited, Kansas City, Mo., Sep. 7, 2004, 2005.
El Naggar, M.H. and M.J.V. Baldinelli, Interpretation of
Axial Statnamic Load Test Using an Automatic Signal
Matching Technique, Canadian Geotechnical Journal,
Vol. 37, 2000, pp. 927942.
Engineering Manual EM 1110-1-2908, Rock Foundations,
U.S. Army Corps of Engineers, Washington, D.C., 1994.
Fleming, W.G.K., A.J. Weltman, M.F. Randolph, and W.K.
Elson, Piling Engineering, John Wiley & Sons, Inc., New
York, N.Y., 1992.
Frantzen, J. and W.F. Stratton, Final Reportp-y Curve
Data for Laterally Loaded Piles in Shale and Sandstone,
Report FHWA-KS-87-2, Kansas Department of Transportation, Topeka, 1987.
Gabr, M.A. and R.H. Borden, LTBASE: A Computer Program for the Analysis of Laterally Loaded Piers
Including Base and Slope Effects, Transportation
Research Record 1169, Transportation Research Board,
National Research Council, Washington, D.C., 1988,
pp. 8393.
Gabr, M.A., R.H. Borden, K.H. Cho, S. Clark, and J.B. Nixon,
P-y Curves for Laterally Loaded Drilled Shafts Embedded
in Weathered Rock, Report FHWA/NC/2002/08, North
Carolina State University, Raleigh, 2002, 289 pp.
Gardner, W.S., Design of Drilled Piers in the Atlantic
Piedmont, Geotechnical Special Publication No. 9:
Foundations and Excavations in Decomposed Rock of the
Piedmont Province, 1987, pp. 6286.
Geophysical Exploration for Engineering and Environmental Investigations, Engineering Manual EM 1110-1-1802,
U.S. Army Corps of Engineers, Washington, D.C., 1995.
Georgiadis, M. and A.P. Michalopoulos, Dilatometer Tests
for the Design of Grouted Piles in Rock, Proceedings, In

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

106

Situ 86, Use of In Situ Tests in Geotechnical Engineering, Blacksburg, Va., 1986, pp. 560568.
Geotechnical Design Manual, Washington State Department
of Transportation, Olympia, 2005 [Online]. Available:
http://www.wsdot.wa.gov/fasc/EngineeringPublications/
Manuals/2005GDM/GDM.htm [Oct. 2005].
Geotechnical Investigations, Engineering Manual EM
1110-1-1804, U.S. Army Corps of Engineers, Washington,
D.C., 2001.
Geotechnical Manual, Texas Department of Transportation,
Austin, 2000 [Online]. Available: http://manuals.dot.state.
tx.us/dynaweb/colbridg/geo [June 2005].
Goodman, R.E., Methods of Geological Engineering in Discontinuous Rock, West Publishing Company, St. Paul,
Minn., 1976.
Goodman, R.E., Introduction to Rock Mechanics, John Wiley
& Sons, New York, N.Y., 1980, 478 pp.
Goodman, R.E. and G.-H. Shi, Block Theory and Its Application to Rock Engineering, PrenticeHall, Englewood
Cliffs, N.J., 338 pp. 1985.
Greer, D.M. and W.S. Gardner, Construction of Drilled Pier
Foundations, John Wiley & Sons, New York, N.Y., 1986,
246 pp.
Gunnink, B. and C. Kiehne, Capacity of Drilled Shafts in
Burlington Limestone, Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 128, No. 7, 2002,
pp. 539545.
Harris, D.E. and P.W. Mayne, Axial Compression Behavior of Two Drilled Shafts in Piedmont Residual Soils,
Proceedings, International Conference on Design and
Construction of Deep Foundations, Vol. 2, Orlando, Fla.,
1994, pp. 352367.
Hassan, K.M. and M.W. ONeill, Side Load-Transfer
Mechanisms in Drilled Shafts in Soft Argillaceous Rock,
Journal of Geotechnical and Geoenvironmental Engineering, Vol. 123, No. 2, 1997, pp. 145152.
Hassan, K.M., M.W. ONeill, S.A. Sheikh, and C.D. Ealy,
Design Method for Drilled Shafts in Soft Argillaceous
Rock, Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 123, No. 3, 1997, pp. 272280.
Hetenyi, M., Beams on Elastic Foundation, The University
of Michigan Press, Ann Arbor, Mich., 1946.
Heuze, F.E., Scale Effects in the Determination of Rock
Mass Strength and Deformability, Journal of Rock
Mechanics, Vol. 12, No. 34, 1980, pp. 167192.
Heuze, F.E., Suggested Method for Estimating the In-Situ
Modulus of Deformation of Rock Using the NX-Borehole
Jack, Geotechnical Testing Journal, Vol. 7, No. 4, 1984,
pp. 205210.
Hiltunen, D.R. and M.J.S. Roth, Investigation of Bridge
Foundation Sites in Karst Terrane Via Multi-Electrode
Resistivity, Proceedings, Geotechnical and Geophysical
Site Characterization, A.V. da Fonseca and P.W. Mayne,
Eds., Millpress Science Publishers, Rotterdam, The
Netherlands, 2004, pp. 483490.
Hirany, A. and F.H. Kulhawy, Conduct and Interpretation of
Load Tests on Drilled Shaft Foundations, Volume 1:

Detailed Guidelines, Report EL-5915, Electric Power


Research Institute, Palo Alto, Calif., 1988, 374 pp.
Hoek, E. and E.T. Brown, Empirical Strength Criterion for
Rock Masses, Journal of Geotechnical Engineering,
Vol. 106, No. GT9, 1980, pp. 10131035.
Hoek, E. and E.T. Brown, Practical Estimates of Rock Mass
Strength, International Journal of Rock Mechanics and
Mineral Science, Vol. 34, No. 8, 1997, pp. 11651180.
Hoek, E. and E.T. Brown, The HoekBrown Failure
CriterionA 1988 Update, Rock Engineering for Underground Excavations, Proceedings 15th Canadian Rock
Mechanics Symposium, Toronto, ON, Canada, 1988,
pp. 3138.
Hoek, E., P.K. Kaiser, and W.F. Bawden, Support of Underground Excavations in Hard Rock, A.A. Balkema, Rotterdam, The Netherlands, 1995, 215 pp.
Hoek, E., C. Carranza-Torres, and B. Corkum, HoekBrown
Failure Criterion2002 Edition, Proceedings, North
American Rock Mechanics Society Meeting, July 810,
2002, Toronto, ON, Canada.
Hoit, M., C. Hays, and M.C. McVay, The Florida Pier
Analysis ProgramMethods and Models for Pier Analysis and Design, Transportation Research Record 1569,
Transportation Research Board, National Research
Council, Washington, D.C., 1997, pp. 18.
Horvath, R.G., Field Load Test Data on Concrete-to-Rock
Bond Strength for Drilled Pier Foundations, Department
of Civil Engineering, University of Toronto Publication
78/07, ON, Canada, 1978.
Horvath, R.G., Drilled Piers Socketed into Weak Rock
Methods of Improving Performance, Ph.D. thesis,
Department of Civil Engineering, University of Toronto,
ON, Canada, 1982.
Horvath, R.G. and T.C. Kenney, Shaft Resistance of Rock
Socketed Drilled Piers, Proceedings, Symposium on
Deep Foundations, ASCE, New York, N.Y., 1979, pp.
182214.
Horvath, R.G., T.C. Kenney, and P. Kozicki, Methods of
Improving the Performance of Drilled Piers in Weak
Rock, Canadian Geotechnical Journal, Vol. 20, 1983,
pp. 758772.
Horvath, R.G., P. Bermingham, and P. Middendorp, The
Equilibrium Point Method of Analysis for Statnamic Loading Test with Supporting Case Histories, Proceedings,
Deep Foundations Conference, Pittsburgh, Pa., 1993.
Hussein, M., B. Robinson, and G. Likens, Application of a
Simplified Dynamic Load Testing Method for Cast-inPlace Piles, Geotechnical Special Publication No. 124,
Geo-Support 2004, 2004, pp. 110121.
Johnston, I.W., T.S.K. Lam, and A.F. Williams, Constant
Normal Stiffness Direct Shear Testing for Socketed Pile
Design in Weak Rock, Geotechnique, Vol. 37, No. 1,
1987, pp. 8389.
Johnston, I.W. and T.S.K. Lam, Shear Behavior of Regular
Triangular Concrete/Rock JointsAnalysis, Journal of
Geotechnical Engineering, Vol. 115, No. 5, 1989, pp.
711727.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

107

Knott, D.L., L.F. Rojas-Gonzales, and F.B. Newman, Current Foundation Engineering Practice for Structures in
Karst Areas, Report FHWA-PA-91-007+90-12, Federal
Highway Administration and Pennsylvania Department
of Transportation, Harrisburg, Pa., 1993.
Koutsoftas, D.C., Caisson Socketed in Sound Mica Schist,
Journal of Geotechnical Engineering, Vol. 107, No. GT6,
1981, pp. 743757.
Kulhawy, F.H., Geomechanical Model for Rock Foundation Settlement, Journal of the Geotechnical Engineering Division, Vol. 104, No. GT2, 1978, pp. 211227.
Kulhawy, F.H. and J.P. Carter, Settlement and Bearing
Capacity of Foundations on Rock Masses, In Engineering
in Rock Masses, F.G. Bell, Ed., ButterworthHeinemann,
Oxford, England, 1992a, pp. 231245.
Kulhawy, F.H. and J.P. Carter, Socketed Foundations in
Rock Masses, In Engineering in Rock Masses, F.G. Bell,
Ed., ButterworthHeinemann, Oxford, England, 1992b,
pp. 509529.
Kulhawy, F.H. and R.E. Goodman, Design of Foundations
on Discontinuous Rock, Proceedings, International
Conference on Structural Foundations on Rock, Vol. 1,
Sydney, Australia, 1980, pp. 209220.
Kulhawy, F.H., C.H. Trautmann, and T.D. ORourke, The
Soil-Rock Boundary: What Is It and Where Is It? Geotechnical Special Publication No. 28: Detection of and
Construction at the Soil/Rock Interface, 1991, pp. 115.
Kulhawy, F.H. and K.-K. Phoon, Drilled Shaft Side Resistance in Clay Soil to Rock, Geotechnical Special
Publication No. 38: Design and Performance of Deep
Foundations, 1993, pp. 172183.
Kulhawy, F.H. and W.A. Prakoso, Discussion of End Bearing Capacity of Drilled Shafts in Rock, Journal of
Geotechnical and Geoenvironmental Engineering, Vol.
125, No. 12, 1998, pp. 11061109.
Kulhawy, F.H., W.A. Prakoso, and S.O. Akbas, Evaluation
of Capacity of Rock Foundation Sockets, Alaska Rocks
2005, Proceedings, 40th U.S. Symposium on Rock
Mechanics, G. Chen, S. Huang, W. Zhou, and J. Tinucci,
Eds., Anchorage, Alaska, June 2005, 8 pp. (CD-ROM).
Ladanyi, B., Friction and End Bearing Tests on Bedrock
for High Capacity Rock Socket Design: Discussion, Canadian Geotechnical Journal, Vol. 14, 1977, pp. 153155.
LaFronz, N.J., D.E. Peterson, R.D. Turton, and S. Anderson,
Geologic Characterization for Bridge Foundations, Colorado River Bridge, Hoover Dam Bypass Project, 54th
Highway Geology Symposium, Burlington, Vt., 2003.
Lai, P., Determination of Design Skin Friction for Drilled
Shafts Socketed in the Florida Limestone, Notes of
Florida DOT Design Conference, Florida Department of
Transportation, Tallahassee, 1998, pp. 140146.
Littlechild, B.D., S.J. Hill, I. Statham, G.D. Plumbridge, and
S.C. Lee, Determination of Rock Mass Modulus for
Foundation Design, Geotechnical Special Publication
No. 97: Innovations and Applications in Geotechnical
Site Characterization, P.W. Mayne and R. Hryciw, Eds.,
2000, pp. 213228.

LRFD Bridge Design Specifications, 3rd ed., American


Association of State Highway and Transportation Officials, Washington, D.C., 2004.
LRFD Bridge Design Specifications, 2006 Interim, Section
10, Foundations, American Association of State Highway
and Transportation Officials, Washington, D.C., 2005.
Manual on Subsurface Investigations, American Association
of State Highway and Transportation Officials, Washington, D.C., 1988.
Marinos, P. and E. Hoek, GSI: A Geologically Friendly
Tool for Rock Mass Strength Estimation, Proceedings,
Geo-Engineering 2000, International Conference on
Geotechnical and Geological Engineering, Melbourne,
Australia, 2000, pp. 14221440.
Matlock, H., Correlations for Design of Laterally Loaded
Piles in Soft Clay, Proceedings, 2nd Offshore Technology Conference, Vol. 1, 1970, pp. 577594.
Mattes, N.S. and H.G. Poulos, Settlement of Single Compressible Pile, Journal of the Soil Mechanics and Foundations Division, Vol. 95, No. SM1, 1969, pp. 189207.
Mayne, P.W., B. Christopher, R. Berg, and J. DeJong,
Manual on Subsurface Investigations, Report FHWA
NHI-01-031, National Highway Institute, Federal
Highway Administration, Washington, D.C., 2001, 301
pp.
McVay, M.C., F.C. Townsend, and R.C. Williams, Design
of Socketed Drilled Shafts in Limestone, Journal of
Geotechnical Engineering, Vol. 118, No. 10, 1992,
pp. 12261237.
McVay, M.C. and L. Niraula, Development of P-Y Curves
for Large Diameter Piles/Drilled Shafts in Limestone for
FBPIER, Final Report to Florida Department of Transportation, Tallahassee, 2004, 158 pp.
Mullins, G., C.L. Lewis, and M.D. Justason, Advancements
in Statnamic Data Regression Techniques, Geotechnical
Special Publication No. 116: Deep Foundations 2002: An
International Perspective on Theory, Design, Construction, and Performance, Vol. 2, 2002, pp. 915930.
Mullins, G., D. Winters, and S. Dapp, Predicting End
Bearing Capacity of Post-Grouted Drilled Shaft in Cohesionless Soils, Journal of Geotechnical and Geoenvironmental Engineering, Vol. 132, No. 5, 2006, pp. 478487.
Murphy, W.L., Geotechnical Descriptions of Rock and
Rock Masses, Technical Report GL-85-3, U.S. Army
Corps of Engineers Waterways Experiment Station,
Vicksburg, Miss., 1985.
ONeill, M.W., F.C. Townsend, K.H. Hassan, A. Buller, and
P.S. Chan, Load Transfer for Drilled Shafts in Intermediate Geomaterials, Report FHWA-RD-95-171, Federal
Highway Administration, McLean, Va., 1996, 184 pp.
ONeill, M.W., D.A. Brown, F.C. Townsend, and N. Abar,
Innovative Load Testing of Deep Foundations, Transportation Research Record 1569, Transportation
Research Board, National Research Council, Washington, D.C., 1997, pp. 1725.
ONeill, M.W. and L.C. Reese, Drilled Shafts: Construction
Procedures and Design Methods, Report FHWA-IF-99-

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

108

025, Federal Highway Administration, Washington,


D.C., 1999, 758 pp.
Ooi, L.H. and J.P. Carter, Direct Shear Behavior of ConcreteSandstone Interfaces, Proceedings, 6th International Conference on Rock Mechanics, Montreal, ON,
Canada, 1987, pp. 467470.
Osterberg, J.O., The Osterberg Cell for Load Testing Drilled
Shafts and Driven Piles: Final Report, Report FHWASA-94-035, Federal Highway Administration, Washington, D.C., 1995, 92 pp.
Osterberg, J.O. and S.A. Gill, Load Transfer Mechanism for
Piers Socketed in Hard Soils or Rock, Proceedings, 9th
Canadian Symposium on Rock Mechanics, Montreal, ON,
Canada, 1973, pp. 235262.
Paikowsky, S.G., et al., NCHRP Report 507: Load and
Resistance Factor Design (LRFD) for Deep Foundations,
Transportation Research Board, National Research Council, Washington, D.C., 2004a, 85 pp.
Paikowsky, S.G., et al., Innovative Load Testing Systems,
Draft Final Report NCHRP 21-08, Transportation
Research Board, National Research Council, Washington, D.C., 2004b.
Paikowsky, S.G., I. Klar, and L.R. Chernauskas, Performance Evaluation of Continuous Flight Auger (CFA) vs.
Bentonite Slurry Drilled Shafts Utilizing Drop Weight
Testing, Geotechnical Special Publication No. 124:
Geo-Support 2004, J.P. Turner and P.W. Mayne, Eds.,
2004c, pp. 637652.
Pells, P.J.N. and R.M. Turner, Elastic Solutions for the
Design and Analysis of Rock-Socketed Piles, Canadian
Geotechnical Journal, Vol. 16, 1979, pp. 481487.
Pells, P.J.N., R.K. Rowe, and R.M. Turner, An Experimental Investigation into Side Shear for Socketed Piles in
Sandstone, Proceedings, International Conference on
Structural Foundations on Rock, Vol. 1, Sydney, Australia, 1980, pp. 291302.
Poulos, H.G., Behavior of Laterally Loaded Piles: I-Single
Piles, Journal of the Soil Mechanics and Foundations
Division, Vol. 97, No. SM5, 1971, pp. 711731.
Poulos, H.G., Behavior of Laterally Loaded Piles: IIISocketed Piles, Journal of the Soil Mechanics and Foundations Division, Vol. 98, No. SM4, 1972, pp. 341360.
Radbruch-Hall, D.H., K. Edwards, and R.H. Batson, Experimental EngineeringGeologic and Environmental
Geologic Maps of the Conterminous United States, U.S.
Geological Survey Bulletin 1610, U.S. Government Printing Office, Washington, D.C., 1987, 7 pp.
Randolph, M.F., The Response of Flexible Piles to Lateral
Loading, Geotechnique, Vol. 31, No. 2, 1981, pp.
247259.
Randolph, M.F. and C.P. Wroth, Analysis and Deformation
of Vertically Loaded Piles, Journal of the Geotechnical
Engineering Division, Vol. 104, No. GT12, 1978, pp.
14651488.
Reese, L.C., Handbook on Design of Piles and Drilled Shafts
Under Lateral Loading, Report FHWA-IP-84-11, Federal
Highway Administration, Washington, D.C., 1984, 386 pp.

Reese, L.C., Analysis of Laterally Loaded Piles in Weak


Rock, Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 123, No. 11, 1997, pp. 10101017.
Reese, L.C. and R.C. Welch, Lateral Loading of Deep
Foundations in Stiff Clay, Journal of the Geotechnical
Engineering Division, Vol. 101, No. 7, 1975, pp.
633649.
Rocha, M., A. Silveirio, F.P. Rodrigues, A. Silverio, and A.
Ferreira, Characterization of the Deformability of Rock
Masses by Dilatometer Tests, Proceedings, 2nd Congress of the International Society of Rock Mechanics,
Belgrade, Yugoslavia, Vol. 1, 1970, pp. 509516.
Rock Slopes: Design, Excavation, Stabilization, Circular
FHWA-TS-89-045, Federal Highway Administration,
Washington, D.C., 1989.
Rock Testing Handbook, Technical Information Center, U.S.
Army Corps of Engineers Waterways Experiment Station, Vicksburg, Miss., 1993.
Rosenburg, P. and N.L. Journeaux, Friction and End Bearing Tests on Bedrock for High Capacity Socket Design,
Canadian Geotechnical Journal, Vol. 13, 1976, pp.
324333.
Rowe, R.K. and H.H. Armitage, The Design of Piles Socketed into Weak Rock, Report GEOT-11-84, University
of Western Ontario, London, ON, Canada, 1984, 365 pp.
Rowe, R.K. and H.H. Armitage, Theoretical Solutions for
Axial Deformation of Drilled Shafts in Rock, Canadian
Geotechnical Journal, Vol. 24, 1987a, pp. 114125.
Rowe, R.K. and H.H. Armitage, A Design Method for
Drilled Piers in Soft Rock, Canadian Geotechnical Journal, Vol. 24, 1987b, pp. 126142.
Rowe, R.K. and P.J.N. Pells, A Theoretical Study of PileRock Socket Behaviour, Proceedings, International
Conference on Structural Foundations on Rock, Vol. 1,
Sydney, Australia, 1980, pp. 253264.
Sabatini, P.J., R.C. Bachus, P.W. Mayne, J.A. Schneider, and
T.E. Zettler, Evaluation of Soil and Rock Properties,
Geotechnical Engineering Circular No. 5, Federal Highway Administration, Washington, D.C., 2002, 385 pp.
Schmertmann, J.H., J.A. Hayes, T. Molnit, and J.O. Osterberg, O-Cell Testing Case Histories Demonstrate the
Importance of Bored Pile (Drilled Shaft) Construction
Technique, Proceedings, Fourth International Conference on Case Histories in Geotechnical Engineering,
St. Louis, Mo., 1998, pp. 11031115.
Schwartz, S.A., Drilled Piers in the PiedmontMinimizing
ContractorEngineerOwner Conflicts, Geotechnical
Special Publication No. 9: Foundations and Excavations
in Decomposed Rock of the Piedmont Province, ASCE,
New York, N.Y., 1987, pp. 87102.
Seidel, J.P. and C.M. Haberfield, A New Approach to the
Prediction of Drilled Pier Performance in Rock, Proceedings, International Conference on Design and
Construction of Deep Foundations, Vol. 2, Orlando, Fla.,
Federal Highway Administration, Washington, D.C.,
1994, pp. 556570.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

109

Seidel, J.P. and B. Collingwood, A New Socket Roughness


Factor for Prediction of Rock Socket Shaft Resistance,
Canadian Geotechnical Journal, Vol. 38, 2001, pp.
138153.
Serafim, J.L. and J.P. Pereira, Considerations of the Geomechanics Classification of Bieniawski, Proceedings,
International Symposium on Engineering Geology and
Underground Construction, Lisbon, Portugal, 1983, pp.
11331144.
Sirles, P.C., NCHRP Synthesis 357: Use of Geophysics for
Transportation Projects, Transportation Research
Board, National Research Council, Washington, D.C.,
2006, 108 pp.
Sowers, G.G., Foundation Bearing in Weathered Rock,
Rock Engineering for Foundations and Slopes, ASCE,
New York, N.Y., 1976, pp. 3242.
Sowers, G.F., Introductory Soil Mechanics and Foundations,
4th ed., Macmillan Publishing Co., Inc., New York, N.Y.,
1979, 621 pp.
Sowers, G.F., Correction and Protection in Limestone Terrain, Proceedings, First Multi-Disciplinary Conference
on Sinkholes, B. Peck, Ed., Balkema, Rotterdam, The
Netherlands, 1994, pp. 373378.
Standard Specifications for Highway Bridges, 16th ed.,
American Association of State Highway and Transportation Officials, Washington, D.C., 1996.
Standard Specifications for Transportation Materials and
Methods of Sampling and Testing, Part II, Tests, American Association of State Highway and Transportation Officials, Washington, D.C., 1992.
Sun, K., Laterally Loaded Piles in Elastic Media, Journal
of Geotechnical Engineering, Vol. 120, No. 8, 1994, pp.
13241344.
Tang, Q., E.C. Drumm, and R.M. Bennett, Response of
Drilled Shaft Foundations in Karst During Construction
Loading, Proceedings, International Conference on
Design and Construction of Deep Foundations, Vol. 3,
Orlando, Fla., Federal Highway Administration, Washington, D.C., 1994, pp. 12961309.
Technical Manual for LPILEPLUS 5.0 for Windows, Ensoft,
Inc., Austin, Tex., 2004.
Terzaghi, R., Sources of Error in Joint Surveys, Geotechnique, Vol. 15, 1965, p. 287.
To, A.C., H. Ernst, and H.H. Einstein, Lateral Load Capacity of Drilled Shafts in Jointed Rock, Journal of Geotechnical and Geoenvironmental Engineering, Vol. 129,
No. 8, 2003, pp. 711726.
Turner, J.P., E. Sandberg, and N.N.S. Chou, Side Resistance of Drilled Shafts in the Denver and Pierre Formations, Geotechnical Special Publication No. 38: Design
and Performance of Deep Foundations, 1993, pp.
245259.
Varnes, D.J., The Logic of Geologic Maps, with Reference
to Their Interpretation and Use for Engineering Pur-

poses, USGS Professional Paper 837, U.S. Government


Printing Office, Washington, D.C., 1974.
Verruijt, A. and A.P. Kooijman, Laterally Loaded Piles in a
Layered Elastic Medium, Geotechnique, Vol. 39, No. 1,
1989, pp. 3946.
Vesic, A.S., NCHRP Synthesis 42: Design of Pile Foundations, Transportation Research Board, National Research
Council, Washington, D.C., 1977, 68 pp.
Viskne, A., Evaluation of In Situ Shear Wave Velocity Measurement Techniques, REC-ERC-76-6, Division of Design, Engineering and Research Center, U.S. Dept. of the
Interior, Bureau of Reclamation, Denver, Colo., 1976.
Wang, S.-T. and L.C. Reese, Analysis of Piles Under Lateral LoadComputer Program COM624P for the Microcomputer, Report FHWA-SA-91-002, Federal Highway
Administration, Washington, D.C., 1991, 229 pp.
Williams, A.F., I.W. Johnston, and I.B. Donald, The Design
of Sockets in Weak Rock, Proceedings, International
Conference on Structural Foundations on Rock, Vol. 1,
Sydney, Australia, 1980, pp. 327347.
Williams, A.F. and P.J.N. Pells, Side Resistance Rock
Sockets in Sandstone, Mudstone, and Shale, Canadian
Geotechnical Journal, Vol. 18, 1981, pp. 502513.
Williams, R., D. Burnett, and J. Savidge, Participant WorkbookDrilled Shaft Foundation Inspection, Publication
FHWA NHI-03-018, U.S. Federal Highway Administration, Washington, D.C., 2002, 776 pp.
Wyllie, D.C., Foundations on Rock, 2nd ed., E&FN Spon,
New York, N.Y., 1999, 401 pp.
Wyllie, D.C. and N.I. Norrish, Rock Strength Properties and
Their Measurement, In TRB Special Report 247: LandslidesInvestigation and Mitigation, A.K. Turner and L.R.
Shuster, Eds., Transportation Research Board, National
Research Council, Washington, D.C., 1996, pp. 372390.
Yang, K., Analysis of Laterally Loaded Drilled Shafts in
Rock, Ph.D. dissertation, University of Akron, Akron,
Ohio, 2006, 291 pp.
Yang, M.Z., M.Z. Islam, E.C. Drumm, and G. Zuo, Side
Resistance of Drilled Shaft Socketed into Wissahickon
Mica Schist, Geotechnical Special Publication No. 124:
Geo-Support 2004, J.P. Turner and P.W. Mayne, Eds.,
2004, pp. 765777.
Zhan, C. and J.-H. Yin, Field Static Load Tests on Drilled
Shaft Founded on or Socketed into Rock, Canadian Geotechnical Journal, Vol. 37, 2000, pp. 12831294.
Zhang, L. and H.H. Einstein, End Bearing Capacity of
Drilled Shafts in Rock, Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 124, No. 7, 1998,
pp. 574584.
Zhang, L., H. Ernst, and H.H. Einstein, Nonlinear Analysis
of Laterally Loaded Rock-Socketed Shafts, Journal of
Geotechnical and Geoenvironmental Engineering, Vol.
126, No. 11, 2000, pp. 955968.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

110

APPENDIX A
Survey Respondents

Alabama
Arizona
Arkansas
California
Colorado
Connecticut
Florida
Georgia
Hawaii
Idaho
Illinois
Iowa
Kansas
Kentucky
Massachusetts
Michigan
Minnesota

Missouri
Montana
New Brunswick
New Hampshire
New Jersey
New Mexico
North Carolina
Oregon
Puerto Rico
South Carolina
South Dakota
Tennessee
Texas
Utah
Vermont
Washington

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

111

APPENDIX B
Survey Questionnaire and Responses

The survey questionnaire is presented in the following pages. Responses to each question are summarized below each question. Some agencies did not respond to every question.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

112

QUESTIONNAIRE
NCHRP TOPIC 36-12
USE OF ROCK-SOCKETED DRILLED SHAFTS
FOR HIGHWAY STRUCTURE FOUNDAT IONS

Background and Purpose


Drilled shafts socketed into rock are widely used as highway bridge foundations and can provide high load capacity
while controlling displacements when designed and constructed appropriately. However, several challenges for
design engineers have been identified in the use of shafts socketed into rock and intermediate geomaterials. These
can be grouped into three categories:

Geotechnical characterization of the rock or intermediate geomaterial


Analysis and design for axial loading
Analysis and design for lateral loading.

The purpose of Synthesis Topic 36-12 is to gather information on how these issues have been addressed in the
design of highway structures. To accomplish the objective, there will be a literature review, survey of bridge owners
from state departments of transportation (DOTs) and toll authorities, and interviews.
This questionnaire is designed to be completed by the state DOT Geotechnical Engineer, assuming that individual
has the most knowledge regarding the issues identified above. However, it is recognized that practice varies
between states and that other branches within a state DOT may have considerable involvement in drilled shaft
engineering. In particular, structural (bridge) engineers responsible for superstructure design may also be involved
in foundation design. Therefore, it is recommended that Part V of this questionnaire (Structural Analysis) be
completed by the state Bridge Engineer. In addition, it is recommended that Part IV (Design for Axial and Lateral
Load) be reviewed by the state Bridge Engineer.
Several questions refer to intermediate geomaterials (IGM) as distinguished from rock. For purposes of this survey,
these two materials are defined as follows:
IGM

cohesive earth material with unconfined compressive strength between 0.5


MPa and 5.0 MPa (5 to 50 tsf) or cohesionless material with SPT N-value
(N60) greater than 50.

Rock

highly cemented geomaterial with unconfined compressive strength


greater than 5.0 MPa (50 tsf).

Part I: Respondent Information


Geotechnical Engineer
Name:
Title:
Agency:
Address:
City:

State:

Zip:

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

113

Phone:

Fax:

e-mail:

City:

State:

Zip:

Phone:

Fax:

e-mail:

Structural or Bridge Engineer


Name:
Title:
Agency:
Address:

Please return the completed questionnaire to:


John P. Turner
Professor, Civil & Architectural Engineering
Department 3295
University of Wyoming
1000 E. University Ave.
Laramie, WY 82071

Phone: 307-766-4265
Fax:
307-766-2221
e-mail: turner@uwyo.edu

After completing the survey, if there are issues pertaining to rock-socketed drilled shafts that you believe are
important but that are not addressed adequately by the questionnaire, please feel free to contact the author directly.

Part II: Defining the Use of Rock-Socketed Drilled Shafts by Your Agency
1.

On approximately how many projects per year (average) does your agency deal with drilled shaft
foundations socketed into rock or IGMs?
None*

(1)

New Brunswick

110

(19)

AZ, AL, AR, CT, HI, ID, IL, KY, ME, MI, MN, NH, NJ, NM, SD, TN,
UT, VT, WA

1020

(7)

GA, IA, MO, MT, OR, PR, SC

More than 20 per year

CA, FL, KS, MA, NC, TX

* If you answered None, skip to Question 5


2.

Please indicate the types of rock or intermediate geomaterials that your agency has dealt with when using
rock-sockets. (Check all that apply.)
Igneous rock types
Granite
Rhyolite
Obsidian
Diorite

Sedimentary rock types


13
5
0
7

Conglomerate
Sandstone
Mudstone/Shale
Limestone

Metamorphic rock types


13
25
24
21

Slate
Phyllite
Schist
Amphibolite

Copyright National Academy of Sciences. All rights reserved.

8
5
12
3

Rock-Socketed Shafts for Highway Structure Foundations

114

Andesite
Gabbro
Basalt
Diabase
Peridotite
Other (describe)

7
5
11
4
1
0

Dolomite
15
Chalk
4
Other (describe) 3

Gneiss
Marble
Quartzite
Serpentinite
Other (describe)

12
2
6
3
2

Intermediate Geomaterials:
claystones, siltstones,
uncemented sandstones
New Jersey
v. dense sands with N > 50
New Hampshire glacial till
New Mexico
Santa Fe Formation
Colorado
(N > 75), of the Rio Grand
Florida
Rift (indurated, cemented,
Georgia
sands, silts, clay)
Illinois
North Carolina weathered rock
Oregon
very soft mudstones,
Iowa
highly weathered
Kentucky
sandstones, weakly
Massachusetts
cemented conglomerates
Michigan
Texas
clay/shales
Minnesota
Utah:
weak shales and
mudstones
Missouri
Vermont:
glacial till
Washington
Has significant deposits of glacial origin. Many have been overridden and overconsolidated by
continental glaciation turning them into IGMs by the definition on page 1. A figure, which can be
accessed at http://www.dnr.wa.gov/geology/pdf/ri33.pdf contains the unit descriptions. The first four
units are encountered in 75% of our shafts.
Alabama
Arkansas
California

clay-shale
hardclay (8 tsf)
mudstone, sandstone, siltstone,
phyllite, slate, and weathered rock
claystone, siltstone, weakly cemented sandstone
weathered limestone
partially weathered rock
weathered limestone, hard clay/shale,
cemented sand/sandstone
shale, siltstone, sandstone, limestone, dolomite
weathered shale
till
soft shale, hardpan
noncemented sandstone,
highly weathered granite
softshale

Montana

Other Rock Type Descriptions:


Hawaii:
Kentucky:
Massachusetts:
Michigan:
3.

tuff
interbedded limestone/shale
argillite
iron ore, coal

Minnesota:
New Mexico:
North Carolina:
Oregon:

meta-graywacke
gypsum
partially cemented rock
diatomaceous siltstone

Indicate the range of rock-socket diameters and lengths used on your agency's projects.

Range of socket diameters:


Alabama
(3.512 ft)
Arizona
(26 ft)
Arkansas
(49 ft)
California
(24 ft)
Connecticut
(2.57.5 ft)
Florida
(312 ft)
Georgia
(4.59.5 ft)
Hawaii
(35 ft)
Idaho
(3.55 ft)
Illinois
(27 ft)
Iowa
(210 ft, usually 3.54.5 ft)
Kansas
(38 ft)
Kentucky
[Typically 47 ft (12 ft max.)]

Missouri
Montana
New Hampshire
New Jersey
New Mexico
North Carolina
Oregon
Puerto Rico
South Carolina
South Dakota
Tennessee
Texas
Utah

(310 ft)
(3.510 ft)
(37.5 ft)
(46 ft)
(2.56 ft)
(3 ft min. to 12 ft max.)
(38 ft)
(34.5 ft)
(28 ft)
(2.510 ft)
(38 ft)
(1.58 ft)
(2.54 ft)

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

115

Maine
Massachusetts
Michigan
Minnesota

(4.57.5 ft)
(Typical 2.54 ft; extreme case10 ft)
(35 ft)
(310 ft)

Ver mont
(510 ft)
Washington
(310 ft, with understanding that 6 ft
and greater may need specialized equipment
or methods)

Range of socket lengths:


Alabama (Geotechnical Foundation Design generally recommends 1 diameter into competent rock. Bridge
structures personnel generally drop this tip elevation to 1.52 diameters.)
Arizona
(1030 ft)
Montana
(550 ft)
Arkansas
(830 ft)
Missouri
(15120 ft)
California
(15 ft to >300 ft)
New Hampshire (330 ft)
Connecticut
(515 ft)
New Jersey
(816 ft)
New Mexico
(630 ft)
Florida
(330 ft)
Georgia
(515 ft)
North Carolina (10120 ft)
Hawaii
(510 ft)
Oregon
(up to 120 ft)
Idaho
(512 ft)
Puerto Rico
(1020 ft)
Illinois
(Typically 510 ft, range 340 ft)
South Carolina (225 ft)
Iowa
(up to 30 ft)
South Dakota
(3090 ft)
Kansas
(4.520 ft)
Tennessee
(1030 ft)
Kentucky
(Typically 615 ft, approx. 30 ft max.)
Texas
(13 shaft diameters)
Maine
(523 ft)
Utah
(210 ft)
Massachusetts (Typically 6 ft; several ft extr. case >50 ft) Vermont
(522 ft)
Michigan
(517 ft)
Washington
(Usually 2 shaft diameters, so 620
Minnesota
(5160 ft)
ft depending on shaft size)
4.

What group or person in your agency has primary responsibility for design of rock-socketed drilled shafts?
(If more than one group within your agency is responsible, please describe briefly the division of tasks
below.)
Geotechnical Branch (30)
AL, AZ, CA, CO, CT, GA, HI, ID, IL, IA, KS, KY, ME, MA, MI, MN, MO, MT, NH, NJ, NM,
NC, OR, SC, SD, TN, TX, UT, VT, WA
Geology Branch (4)
AZ, CA, KS, MN
Bridge Engineering (structural) (20)
AL, AR, CA, CO, CT, GA, ID, IA, KS, KY, MA, MI, MN, MO, NH, OR, SD, TN, UT, VT
Outside Consultant (11)
CT, FL, HI, ID, IL, IA, KY, MA, NB, NM, PR
Other (explain): Iowa (Input from FHWA)
Division of tasks (if applicable):
AL: Geotechnical is responsible for axial capacity. Bridge is responsible for lateral stability.
CT: All our drilled shaft projects to date have been designed by outside consultants. If we were doing the
design in-house, the responsibility would be shared between the geotech and bridge designer.
CO: Geotech provides geotechnical design parameters. Bridge performs design.
FL: AxialGeotechnical; LateralStructural.
GA: GeotechnicalSelection of shafts is recommended foundation type or alternate; bearing pressures, rocksocket length, and tip elevations. BridgeSelection of shaft diameters, lateral analysis, and possible revision
of tip elevations.
KS: Geotechs set base of shaft elevation and recommend side shear and end bearing strengths.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

116

KY: Geotechnical Branch and/or Geotechnical ConsultantGeotech investigation, axial capacities, tip
elevations. Division of Bridge Design and/or Structures ConsultantStructural design and detailing, structure
plans.
MA: GeotechnicalDimensions and capacities based on loadings and soil/rock properties. Structuralrebar,
concrete, connection designs.
MI: Geotechnical characterizes rock formation and determines rock-socket diameter and length. Bridge
Design determines shaft location, shaft loading, and sizes reinforcement.
MN: Geotech determines design bearing capacities and soils and rock properties with consultation with
geology. Structures designs final shaft dimensions.
MO: Geotechnical Office provides design criteria and evaluates shaft design based on materials encountered
and proposed shaft configuration. Bridge Engineering proposes the layout of foundation units and designs the
shaft itself (size, steel configuration, etc.).
NM: Geotechnical Section approves outside consultant designs.
UT: Geotech Branchrock resistance, L-pile; Structuresstructural design.
VT: Geotechnical capacity and lateral analysis is done by the geotechnical branch and structural design is
done by the structures group.
WA: Geotechs assess capacity and settlement and provide p-y input parameters to bridge office. The
structural designer in the bridge office performs the structural design of the shaft assessing shear, moments,
and rebar/concrete requirements. They also perform the seismic design using the geotechs p-y parameters.
5.

In the next 3 years, do you anticipate that the use of rock-socketed drilled shafts by your agency will:
Increase (8)
ID, KS, MA, MO, NB, PR, SC, TN
Remain approximately the same (25)
AL, AZ, AR, CA, CT, FL, GA, HI, IL, IA, KY, ME, MI, MN, MT, NH, NJ, NM, NC, OR, SD,
TX, UT, VT, WA
Decrease (none)

6.

Please add any comments you feel would be useful, pertaining to the use of rock-socketed drilled shafts by
your agency.
CA: Most result in claims due to the requirement to include Differing Site Conditions on all contracts.
IA: Use of drilled shafts has been more frequent in past 23 years (above historic use), but may fall off
again within next 23 years.
KS: Used on high tower lighting and sign structure footings. Used as a contractors option on some
structures. Ease of construction around highway and railway facilities.
MS: Combination of new codes and loadings, issues of scour, and extreme events, are driving the use of
deep and/or rock-socketed shafts.
MO: Use is increasing in part due to more consultant bridge design and MoDOT bridge designers gaining
more experience in shaft design. Shaft design is cost-competitive with driven pile in many cases and
construction in urban areas causes less noise and vibration than driven pile.
New Brunswick: We are beginning to recognize the potential of this type of foundation as an option for
bridge foundations in our province. We currently have two projects in the design stage that will use drilled
shaft foundations. Where we have limited design experience in-house; most of the questions in the survey
are left unanswered. We look forward to reviewing the results as a way to see how other agencies approach
these designs, as we move toward the consideration of drilled shafts as an option in the future.
NH: Emphasize that the design of drilled shafts should include consideration of the drilled shaft
construction methods and constructability issues.
OR: Most of our shafts are socketed into bedrock, either with or without end bearing resistance. Many
times we need rock embedment to resist high lateral loads associated with high seismic loading conditions,
sometimes coupled with soil liquefaction.
SC: Finding some way to equate different rock drilling rigs/equipment capabilities to varying rock
strengths.
SD: 99% of the drilled shafts done in this state are done in shale bedrock.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

117

WA: When we want to have a rock socket of a certain length and recognize that the rock may be variable
in elevation we include the following provision in our shaft special provision. With this special in place,
the contractor can tie the reinforcing cage prior to excavating, excavate to rock, construct the rock socket,
and trim the cage to fit. Excavation to tip elevation, cage placement, and the concrete pour can be complete
in one shift this way. We pay for the steel that is cut off from the bottom of the cage, but feel that it is well
worth the investment by lowering our risk of a blow-in or caving as the shaft does not have to sit open for
days while the cage is tied. When the contract requires a minimum penetration into a bearing layer, as
opposed to a specified shaft tip elevation, and the bearing layer elevation at each shaft cannot be accurately
determined, add subsection 3.05.E as follows: For those shafts with a specified minimum penetration into
the bearing layer and no specified tip elevation the Contractor shall furnish each shaft steel reinforcing bar
cage, including access tubes for cross-hole sonic log testing in accordance with subsection 3.06 of this
Special Provision, 20% longer than specified in the plans. The Contractor shall add the increased length to
the bottom of the cage. The contractor shall trim the shaft steel reinforcing bar cage to the proper length
prior to placing it into the excavation. If trimming the cage is required and access tubes for cross-hole
sonic log testing are attached to the cage, the Contractor shall either shift the access tubes up the cage or cut
the access tubes provided that the cut tube ends are adapted to receive the watertight cap as specified.

Part III: Characterization of Rock or Intermediate Geomaterial (IGM)


7.

Check the methods used by your agency to determine depth to bedrock for the purpose of drilled shaft
foundation engineering.
Standard Penetration Test (SPT) refusal
(22)
AR, CA, CO, FL, GA, HI, IL, IA, KS, MA,
MI, MN, MO, MT, NJ, NM, NC, OR, PR, SC, UT, VT
Cone Penetration Test (CPT) refusal:

(3)

KS, MN, MO

Coring and inspection of core samples


(30)
AL, AZ, AR, CA, CO, CT, FL, GA, HA, ID,
IL, IA, KS, KY, ME, MA, MI, MN, MO, NH, NJ, NM, NC, OR, PR, SC, TN, UT, VT, WA
Geophysical methods (specify)
AZ, CA, ID (seismic refraction), KS, KY (currently using resistivity and microgravity on a project
where we are considering drilled shaft foundations. We have not previously used geophysical
methods for drilled shaft investigations), MT, UT, (seismic refraction), WA
Other (specifyprovide details if possible)
CA (2.25 in diameter terry cone driven to refusal. Please note that no one method is solely relied
upon), IL, IA (laboratory UC or other tests), MN (pressuremeter), NJ (point load strength test),
NM (RQD/RMR/joint orientation/water/polymer slurry only), SD (California retractable plug
sampler to extract samples), TX (Texas Cone Penetrometerfor information see the following
websites: http://txdot-manuals/dynaweb/colbridg/geogo to Chapter 2, Section 4, Field Testing
for Design Chartsgo to Chapter 4, Design Guidelines: http://txdotmanuals/dynaweb/colmates/soi/@Generic__BookView;cs=default;ts=default; go to Section 32
Tex132-E Texas Cone Penetration Test).
8.

How does your agency distinguish between rock, soil, and intermediate geomaterials?
Defined in the same way as stated on page 1 of this questionnaire
(24)
AL, AR, CA, CO,
FL, HI, ID, IL, KS, KY, ME, MI, MN, MO, MT, NM, OR, PR, SC, SD, TN, UT, VT, WA
Other: summarize below
AZ: Typically, we classify as either soil or rock only by the use of test borings.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

118

CT: We generally do not try to quantify IGMs. We may have some glacial tills/weathered
bedrocks overlying a hard bedrock that would be an IGM, but we do not usually spend much time
defining its engineering properties for the design of drilled shafts.
CO: Some very weathered claystone is classified as rock even if it is weaker than IGM, as
described in the background and purpose section above.
GA: Soil-drilled and sampled with earth augers, SPT < 50, drilled shaft bearing pressure < 30
40 ksf; IGM-drilled with earth and/or rock augers, SPT > 50, drilled shaft bearing pressure >
40 ksf, < 75 ksf; rock material below auger refusal sampled with diamond core drilling, drilled
shaft bearing pressure > 75 ksf.
IL: Experience combined with field observation of drilling operation (difficulties, change of
drilling tools, etc.)
IA: Classify as IGM? Rock if of sedimentary rock geologic origin. Classify as soil if of
glacial, alluvial, similar deposition.
KY: We have very few IGMs and if we have them, they are typically a weathered zone of shale in
a transition from residual soil to interbedded limestone and shale. This material is typically
neglected for drilled shaft design.
MA: We have a clear distinction between rocks and soils, based on coring use.
ME:
NH: For classification purposes on test boring logs, differentiation of bedrock vs. IGM or soil
based on geologic interpretation of boring samples. For drilled shaft analysis, would generally use
the definitions on page 1 (e.g., weathered bedrock would be classified on the boring log as
bedrock, but would be analyzed as an IGM).
NJ: Based on the coring results, RQD and recovery, and engineering judgment; e.g., RQD < 30%
may be considered as IGM not sound rock.
NC: Definition of RockSPT and refusalRock is defined as a continuous intact natural
material in which the penetration rate with a rock auger is less than 2 in. (50 mm) per 5 min of
drilling at full crowd force. This definition excludes discontinuous loose natural materials such as
boulders and man-made materials such as concrete, steel, timber, etc.
TX: Our design methodology does not require specific designation of rock, soil, or IGM. Design
is generally based on the strength testing, regardless of material designation.
The following is a list of rock properties that may be required or recommended to apply design methods specified in
the FHWA Drilled Shaft Manual, as well as for other published design methods used for rock-socketed drilled
shafts:
=
=
=
=

qu
RQD
RC
Ecore

unconfined compressive strength (units of F/L2)


Rock Quality Designation
effective stress angle of friction between the rock or IGM and concrete
Young's modulus of rock or IGM core (units of F/L2)

Rock Mass Quality as defined in terms of:


RMR
=
Rock Mass Rating (Bieniawski 1974)
Q
=
Norwegian Geotechnical Institute rating (Barton 1974)
c'i and 'i
=
instantaneous values of cohesion and friction angle for HoekBrown
nonlinear strength criteria of fractured rock masses
9.

For each rock property, check the appropriate box indicating whether your agency determines this property
for rock-socket design and, if so, the method used to determine the property:
qu

Always

Never

Varies

Always: (23) AL, AZ, CT, FL, HI, ID, IL, KS, ME, MA, MI, MN, MO, MT, NJ, NM, OR, SC, SD, TN,
UT, VT, WA
Never: (0)
Varies : (10) AR, CA, CO, GA, IA, KY, NH, NC, PR, TX

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

119

Method:
ASTM D2938 or AASHTO T-226: Uniaxial Compressive Strength
(14)
AL, AZ, IA, KY, MN, MO, NH, NM, OR, SC, TN, TX, UT, VT
Point Load Tests and/or Uniaxial Compression of intact core:
(3)
MA, MI, WA
Maine (ASTM D7012-04) ??
RQD

Always

Never

Va ries

Always: (28) AL, AZ, AR, CA, CO, CT, GA, HI, ID, IL, IA, KS, ME, MA, MI, MN, MO, MT, NH, NJ,
NC, OR, PR, SC, TN, UT, VT, WA
Never: (1) SD
Varies: (3) FL, KY, TX
RC

Always

Never

Va ries

Always: (2) KS, MT


Never: (23) AL, AR, CT, FL, GA, HI, ID, IL, KY, ME, MA, MI, MN, NH, NM, NC, OR, PR, SC, SD,
TX, UT, VT
Varies: (7) AZ, CA, CO, IA, NJ, TN, WA
Method: AL (Relies on charts), AZ (Estimate from AASHTO Manual), IA (theoretical),
WA (Usually use published textbook values based on rock type)
Ecore

Always

Never

Va ries

Always: (5) KS, ME, MN, UT, VT


Never: (11) AL, AR, CA, HI, ID, MT, NM, PR, SC, SD, TX
Varies: (16) AZ, CO, CT, FL, GA, IL, IA, KY, MA, MI, NH, NJ, NC, OR, TN, WA
Method:
AL: Correlation charts between qu and E
IA: Theoretical
KY: Correlation with UC strength
ME: ASTM 7012-04
MA: Goodman, Jack, and tables/charts
MI: Calculated from Ultrasonic Velo city test (ASTM D2845) or approximated from figures and
tables in section 4 of the AASHTO Standard Specs
MN: ASTM D3148
NH: Eintact determined from qu test, then correlated to Ein situ through RMR or other methods
OR: From ASTM D2938 results with measured strains
UT: either unconfined compression test or from AASHTO table
VT: ASTM D3148
WA : Usually use published textbook values based on rock type
RMR

Always

Never

Va ries

Always: (5) MA, NH, NM, TN, UT


Never: (15) AL, AR, FL, HI, IL, KY, ME, MI, MN, MT, NJ, PR, SC, SD, TX
Varies: (12) AZ, CA, CO, CT, GA, ID, IA, KS, NC, OR, VT, WA

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

120

Always

Never

Varies

Always: none
Never: (29) AL, AZ, AR, CA, CO, CT, FL, GA, HI, ID, IL, IA, KY, ME, MA, MI, MN, MT, NH, NJ,
NM, OR, PR, SC, SD, TX, UT, VT, WA
Varies : (3) KS, NC, TN
c'i and 'i

Always

Never

Varies

Always: none
Never: (24) AL, AR, CT, FL, GA, HI, ID, IL, IA, KS, KY, ME, MA, MI, MN, MT, PR, SC, SD, TN,
TX, UT, VT, WA
Varies : (8) AZ, CA, CO, NH, NJ, NM, NC, OR
10.

List below any in situ test methods that are used by your agency to correlate with rock or IGM properties or
to correlate directly to rock-socket design parameters (e.g., side or end bearing resistance).
In Situ Test
Standard Penetration Test (SPT)

Pressuremeter Test (PMT)

Borehole (Goodman) Jack


Dilatometer
TxDOT Cone Penetrometer
11.

State
CA
HI
IL
MO
NH
WA
FL
AZ
CA
MN
OR
MA
CA
TX

Property or Design Parameter


not stated
strength
inch penetration per 100 blows (no property stated)
correlation to qu
side and end bearing resistances
friction angle for IGMs
strength
shear values
rock mass modulus
stiffness (rock mass modulus)
correlation with p-y curves
rock mass modulus
rock mass modulus
correlate to side and tip resistances

Indicate by marking the boxes whether your agency uses any of the following tools for evaluating
characteristics of rock below base elevation:
Coring into the rock below the bottom of the shaft after the excavation to base
elevation is complete; if so, to what depth?
AZ: 3B or minimum of 10 ft
AL: typically 10 ft unless specified otherwise
FL: >10 ft
GA: 6 ft
MO: 10 ft below the bottom of the shaft for end bearing design; not required when designed for side
friction only
MT: 50 ft
NJ: not stated
NM: 3 diameters
NC: 5 ft
TX: at least 5 ft deep or a depth equal to the shaft diameter, whichever is greater.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

121

Coring into rock below bottom elevation prior to excavating the shaft.
(27) AL, AZ, AR, CA, CT, FL, GA, HI, ID, IL, IA, KS, KY, MA, MI, MN, MO, NH, NJ, NM,
NC, OR, PR, TN, UT, VT, WA
Inspection of core holes at the bottom of the socket using feeler rods.
(5) AL, GA, MA, NC, TN
Inspection of core holes using fiberoptic cameras.
(5) AZ, NJ, NM, NC, UT
Other:

CA: Visually inspect drilled hole and cores and/or cuttings that are removed.
IL: Visual inspection and classification of rock core by an experienced geologist.
ME: Camera inspection of rock-socket base and extending borings during design stage to depth
below expected bottom of rock socket.
NC: 10 lb weight, SID camera, or use temporary casing to inspect the base by the engineer or the
contractor.
SC: Corings into rock below shaft bottom during design represent expected rock below base.
UT: Visual inspection; many times, the rebar cage is designed to go to the bottom of the boring (in
shorter shafts)this verifies depth.

12. If your agency has experience in the design and construction of rock sockets with any of the materials listed
below, please check the appropriate box and provide information on test methods (field or laboratory) that
you have used to characterize the material properties
Weak lime rock

(11)

AZ, FL, GA, IA, KS, MN, MO, NC, SC, TX, UT

State: Property; Test Method; Correlations Used


AZ: This applies to all rock types listed below: shear and end bearing; unconfined compressive strengths
and RQD; AASHTO Guidelines
FL: Coring, qu, qt, RQD, Recovery (%), SPT
GA: Coring, RQD, compressive tests, split tensile
IA: Strength, skin friction/end bearing; lab UC on cores, O-cell tests
KS: Core, RQD RMR qu
MN: Strength and stiffness; unconfined compression
MO: Compressive strength; qu on core sample
NC: Typically we core the rock and perform unconfined compression tests
SC: Unconfined; load test
TX: qu, skin friction, point bearing; ASTM, Tex-132-E; TxDOT Geotechnical Manual correlations
UT: Same mentioned in Question 9
Soft shales or marls

(14)

AL, AZ, CA, GA, IA, KS, KY, MI, MO, NM, SC, SD, TX, UT

State: Property; Test Method; Correlations Used


AL: Rock strength, thickness and spacing of discontinuities; unconfined compression testing where
possible, logged by a professional geologist from the cored rock (for all rock types checked)
CA: Unconfined strength; triaxial test
GA: Coring, RQD, compression tests
IA: Strength skin friction/end bearing; lab UC on cores and O-cell tests
KS: As above
KY: Unconfined compression, slake durability index
MI: Unconfined compressive strength; ASTM D2166
MO: Compressive strength; qu on core and correlations with SPT; Texas DOT correlations modified by

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

122

MoDOT
NM: qu/triax shear; AASHTO T296; Alpha method
SC: Triaxial; load tests
SD: Soil strengths; unconfined compression test; skin resistance compared to pull test on steel pk rod
TX: qu, skin friction, point bearing; ASTM, Tex-118-E, Tex-132-E; TxDOT Geotechnical Manual
correlations
UT: Same as mentioned in Question 9
Weathered and highly fractured rock
(20)
AL, AZ, CA, GA, HI, IA, KS, KY, MA, MI, MN, MO, NM, NC, OR, SC, SD, TX, UT, WA
State: Property; Test Method; Correlations Used
AR: Visual observations of rock condition
GA: Coring, RQD, compression tests
HI: Strength; unconfined compression
IA: Strength skin friction/end bearing; lab UC on cores and O-cell tests
KS: As above
KY: We may use Slake Durability Index in shale and sometimes sandstone
MA: RMR/qu
MI: Unconfined compressive strength; Point Load Test ASTM D5731; correlations included in test
procedure
MN: Strength and stiffness; SPT or pressuremeter
MO: RQD
NM: phi'; N60; Mayne and Harris
OR: Shear strength; SPT, judgment based on experience, often treated as very dense granular soil;
Meyerhof or Peck, Hanson, Thornburg
SC: SPT
SD: Soil strengths; unconfined compression test; skin resistance compared to pull test on steel pk rod
TX: qu, skin friction, point bearing; ASTM, Tex-132-E; TxDOT Geotechnical Manual correlations
UT: Same mentioned in Question 9
WA: RQD and unconfined compressive strength; drilling and Point Load
Karst

(9)

AL, AZ, FL, GA, KS, KY, MI, NM, TX

State: Property; Test Method


AZ: Core into rock after the excavation to check for voids
FL: Coring, qu, qt, SPT
GA: Coring, RQD, compressive tests, split tensile
KS: Seismic
KY: Rock Core Recovery
NM: Discontinuities; test pits/seismic shear wave
TX: qu, skin friction, point bearing; TxDOT Geotechnical Manual correlations
Rock with steeply dipping discontinuities

(7)

AL, AR AZ, CA, GA, NM, WA

State: Property; Test Method; Correlations


AR: Visual observations of rock condition
AZ: Down-the-hole camera to check for poorly oriented joint sets
GA: Coring, RQD, compressive tests
NM: Modulus; RMR
WA: RQD and unconfined compressive strength; drilling and Point Load
Interbedded rock with alternating strong and weak strata
AL, AR, AZ, CA, GA, IA, KS, KY, MO, NM, TX

(11)

State: Property; Test Method

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

123

AR: Visual observations of rock condition


GA: Coring, RQD, compressive tests
IA: Strength skin friction/end bearing; lab UC on cores, O-cell tests
KS: Core, qu, RMR, RQD
KY: Unconfined compression, Slake Durability Index
MO: Compressive strength; qu from representative core samples
NM: Side shear/modulus; Rowe and Armitage
TX: qu, skin friction, point bearing; ASTM, Tex-118-E, Tex-132-E
Hard, intact rock
(22)
AL, AZ, CA, CT, GA, HI, ID, IA, KY, ME, MA, MI, MN,
MO, NH, NM, NC, OR, SC, TX, VT, WA
State: Property; Test Method; Correlations
AL:
AZ:
CA:
CT: Unconfined compressive strength
GA: Coring, RQD, compressive tests
HI: Strength; unconfined compression
ID: Unconfined compressive strength; ASTM D2938
IA: Strength skin friction/end bearing; Lab UC on cores O-cell tests
KY: Unconfined compression
ME: qu and E; D 7012-04
MA: qu; point load: Ip; 25 Ip = qu
MI: Unconfined compressive strength; ASTM-C42
MN: Strength and stiffness; unconfined compression test
MO: Compressive strength; qu from core samples
NH: Intact compressive strength; unconfined compression test
NM: qu; RMR
NC:
OR: Unconfined compressive strength; ASTM D2938
SC: Unconfined; FHWA methodology
TX: qu, skin friction, point bearing; ASTM, Tex-132-E; TxDOT Geotechnical Manual correlations
VT: qu; ASTM D2938; FHWA IF-99-025
WA: RQD and unconfined compressive strength; drilling and point load
13.

Identify any other issues pertaining to IGM or rock characterization that you think should be addressed by
the Synthesis.
MO: Limited Osterberg load cell testing has indicated that we significantly overdesign shafts in IGMs
based on compressive strength values from qu testing. We need low cost in situ or other test methods for
obtaining ultimate capacities in IGMs.
NC: NCDOT and the NC State University conducted research to determine p-y curves for soft weathered
rock loaded horizontally.
OR: In Question 9, is anyone actually measuring or estimating the s and m dimensions of the rock
mass for the Carter and Kulhawy equation? Also, is anyone estimating borehole roughness and using the
Horvath (1983) equation? We are not because we have no real way of knowing if this can be accomplished
in the field.
UT: How are states handling discontinuities in design; how are strength values of the discontinuities being
determined, etc.?
WA: In Washington State, our shaft lengths are rarely designed to carry the applied axial loads. Most
shafts have very significant lateral capacity demand owing to earthquake loading. Tip elevations are often
set to meet lateral capacity requirements. Very little information is available on the lateral capacity or
lateral behavior of shafts in IGMs subjected to lateral loads. The effects of group loading in IGMs are also
not well-documented.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

124

Part IV: Design for Axial and Lateral Load


Note: The terms base resistance, tip resistance, and end-bearing resistance are used by various agencies; all
refer to the resistance developed beneath the tip of a deep foundation.
14.

When designing for axial load of rock-socketed shafts, does your agency account for:
Both side and base resistances
(25)
AL, AZ, CA, CO, CT, FL, GA, IA, KS, KY, ME, MI,
MN, MO, MT, NM, NC, OR, PR, SC, TN, TX, UT, VT, WA
Side resistance only

(10)

CA, HI, ID, MA, MN, MO, NH, NJ, SD, TN

Base resistance only

(7)

AR, CA, ID, ME, MA, MO, TN

Comments:
AZ: Rely mainly on side resistance with reduced end bearing.
CA: Depends on anticipated methods of construction.
IL: Depends on the elastic deformation. Generally, not both side and end.
MO: Evaluated on a case-by-case basis. Typically end bearing only in hard rock and side resistance only
in alternating hard and soft rock layers.
NH: Rock-socket lengths typically controlled by lateral load with sufficient geotechnical capacity provided
by side shear only. Would consider using a portion of the end bearing geotechnical capacity in
combination with full side shear, if needed, to avoid extending socket length beyond what may be
needed for lateral loads.
NC: Depends on our design; we might use base or side resistance but most of the time we use both.
OR: Combine side and base resistance only in very ductile rock formations as described in the FHWA
Drilled Shaft Design Manual.
TN: Geotech Section provides parameters for both. Structure designer decides which to use.
UT: In wet conditions we will, many times, discount base resistance; with the current LRFD code we have
been using either side resistance only (most of the time) or base resistance, based on deflection.
15.

For calculating side resistance of rock sockets, please indicate the reference(s) associated with the
method(s) used by your agency (mark all that apply):
O'Neill and Reese (1999) Publication No. FHWA-IF-99-0 25, Drilled Shafts: Construction
Procedures and Design Methods
(26)
AL, AZ, CA, CT, GA, HI, ID, IA, KS, KY, MA, MI, MN, MO, MT, NH, NJ, NM, NC,
OR, PR, SC, TN, UT, VT, WA
Horvath and Kenney (1979)
(8)
CA, HI, ID, IL, ME, OR, SC, TN
Rowe and Armitage (1984)
(1)
NM
Carter and Kulhawy (1988)
(6)
CA, IL, ME, NC, SC, WA
Other (please cite reference or provide a brief description)
(6)
AZ: (AASHTO 2002)
KS: (Results from O-cell tests)
FL: McVay et al. (1992)
NH: (2002 AASHTO bridge code)

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

125

OR (not stated)
TX (TxDOT Design MethodSee Chap. 4 of the Geotechnical Design Manual (website)
16.

What, if any, computer programs are used by your agency for analysis of rock-socket response to axial
loading?
SHAFT (14)
AL, AZ, CA, GA, KS, MN, MT, NJ, NM, NC, OR, PR, SC, UT
FBPIER (6)
CT, FL, MN, NC, TN, VT
ROCKET (0)
CUFAD (0)
Other (name of program) (4)
FL: FB Deep
IL: In-house spreadsheet based on Pells and Turner
KY: In-house spreadsheets
TX: WinCoreTxDOT program for the design of drilled shafts

17.

Specify the range of values used by your agency for either Factor of Safety (FOS) or Resistance Factor (s)
applied to rock-socket ultimate side resistance in design (if applicable, specify by rock type).
AL:
AZ:
CA:
CT:

FOS 3.0
All, FOS 2.5
Hard rock, FOS 2.02.5 or s 1 or 0.75
AASHTO ASD, LFD, or LRFD recommended
values regardless of rock type
FL: By AASHTO LRFD
GA: Weak IGM or hard granite FOS 2.5
HI: Tuff, s 0.65
ID: Igneous (basalt), s 0.550.65
IL: All, FOS 2.5
IA: IGM and Rock, FOS 22.5
KY: All, FOS 2 (if load tested) to 3
MA: All, FOS 22.5 or s 0.550.65
ME: Schist, FOS 2.5

18.

MI: All, FOS 3


NH: All, FOS 2.5
NJ: All rocks, FOS 2
NM: All, FOS 2.0
NC: All, FOS 2.53.00, We reduce the
FOS if we perform a load test
OR: All, FOS 2.5
SC: All, FOS 23, s 0.40.7
SD: Shale, FOS 2.0?
TX: All, FOS 3
UT: All types, s 0.55
VT: All, FOS 2.5
WA: All, FOS 3.0 Static, 1.65 Seismic

For calculating base resistance of rock sockets, please indicate the reference(s) associated with the
method(s) used by your agency (mark all that apply):
O'Neill and Reese (1999) Publication No. FHWA-IF-99-025, Drilled Shafts:
Construction Procedures and Design Methods
(24)
AL, AR, CA, CT, GA, ID, IA, KS, KY, ME, MA, MI, MN, MO, MT, NH, NJ, NM, NC,
OR, PR, UT, VT, WA
Canadian Foundation Engineering Manual (1979)
(3)
CA, IL, ME
Zhang and Einstein (1998)
(2)
ID, MA
Carter and Kulhawy (1988)
(3)
AR, CA, WA
Other (please cite reference or provide a brief description)
AZ: AASHTO 2002
KS: O-cell tests results
KY: Experience and judgment

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

126

NH: 2002 AASHTO Bridge code


NM: Rowe and Armitage
TN: Use Soils and Geology Allowables and Section 4 of AASHTO Specs.
TX: TxDOT Design MethodSee Chap. 4 of the Geotechnical Design Manual (website)
WA: AASHTO LRFD Manual
19.

Specify the range of values used by your agency for either Factor of Safety (FOS) or Resistance Factor (s)
applied to rock-socket ultimate base resistance in design:
AL:
AR:
AZ:
CA:
CT:

FOS 3.0
Sandstone or shale, FOS 2.5
All, FOS 2.5
All, FOS 2.0 or s 1 or 0.75
AASHTO ASD, LFD, or LRFD recommended
values regardless of rock type
FL: Side/end: 0.55; side only: 0.60
GA: Weak IGM or hard granite, FOS 2.5
ID: Igneous (basalt), s 0.5
IL: All, FOS 2.5
IA: IGM and rock, FOS 22.5
KY: All, FOS 2 (if load tested) to 3
MA: All, FOS 22.5 or s 0.5
ME: Schist, FOS 2.5

20.

MI: All, FOS 3


MN: FOS 2.5
NH: All, FOS 2.5
NJ: All rocks, FOS 2
NM: All, FOS 2.0
NC: All, FOS 2.53.0
OR: All, FOS 2.5
SC : All, FOS 23, s 0.40.7
SD : Shale, FOS 2.0?
TN: All, FOS 2.5
TX: All, FOS 2
UT: All types, s 0.5
VT: All, FOS 2.5
WA: All, FOS 3.0 Static, 1.65 Seismic

If you include both side and base resistances in design of rock sockets, explain briefly how you account for
the relative contribution of each to the socket axial resistance
CA: Must determine the amount of each that can be mobilized at our allowable movement at the top of the
pile.
CT: The relative contributions would be based on the computed displacement/strain of the drilled shaft. If
load test data were available, the strain compatibility would be validated or refined based on the
actual test data.
FL: Based on compatibility.
IL: For weak IGM.
IA: Both are typically limited by settlement criteria for both allowable and ultimate loads.
KS: In good hard rock most of the load is stripped off in side shear. In shales, we assume that side shear
and end bearing act together; either O-cell testing at site or extrapolation of previous testing.
KY: Evaluate strain compatibility if O-cell load test is run. If no load test, use higher FS (3).
ME: Two projects were designed in accordance with AASHTO 4.6.5.3, assuming that axial loads are
carried solely by end resistance, as the strains required for full mobilization of both end and side
resistance is incompatible. This design approach was later altered on one project to assume
conservative, simultaneous mobilization of both end and side frictional resistance. That design
approach ignored side resistance in the upper 5 to 15 ft may (based on a minimum required value of
RQD and qu, determined by the geotechnical engineer). For the remainder of the side walls, partial
contribution is assumed, in addition to partial mobilization of full end bearing.
MI: Seek to design socket to have side friction capacity 2.0 to 2.5 times applied load. When end bearing
contribution is added, seek to show FS greater than or equal to 3.0.
NM: Osterberg and Gill (length/modulus ratio).
NC: This assumption will depend on engineering judgment and the method of construction.
OR: According to methods described in FHWA manual; determine the resistance available from the side
and base independently based on a given relative shaft settlement and then add them together.
SC: Assume side fully mobilized and 5% diameter settlement not necessary to mobilize end resistance in
rock.
TN: Geotech Section opinion is that with the rock type and strength we have and using a safety factor of
2.5, a relatively small mobilization of side and end bearing occurs; therefore, it is okay to use a
combination of both. Structures Designer typically uses just one or the other.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

127

TX: TxDOT Design MethodSee Chap. 4 of the Geotechnical Design Manual (website cited above).
WA: See the attached pdf discussing WSDOT procedures for designing drilled shafts in rock and IGMs.
21.

When a bridge is supported on a shallow footing that is supported on a rock-socketed drilled shaft (as
opposed to a mono shaft), does your design procedure account for the contribution of the footing to the
foundation capacity?
Ye s

(none)

No
(25)
AL, AR, CA, CT, FL, GA, HI, IL, IA, KS, KY, MA, MI, MN, MT, NH, NM,
NC, SC, SD, TN, TX, UT, VT, WA
Not applicable

(6)

AZ, ID, ME, NJ, OR, PR

If you answered Yes, please provide a brief description of your analysis to account for the footing
contribution: (none)
22.

For analysis of rock-socketed shafts under lateral loading, please indicate the methods and/or references
associated with methods used by your agency (mark all that apply).
Equivalent Cantilever Method (Davisson 1970)

(5)

KS, MA, NH, NC, SC

Broms Method (Broms 1964)

(5)

KY, MT, OR, SC, TN

p-y method of analysis


(26)
AZ, AR, CA, CT, FL, GA, HI, IL,
IA, KS, KY, MA, MI, MN, NH, NJ, NM, NC, OR, PR, SC, TN, TX, UT, VT, WA
Characteristic Load Method (Duncan et al. 1994)

(none)

Zhang, Ernst, and Einstein (2000) Nonlinear Analysis of Laterally Loaded Rock-Socketed
Shafts
(1)
MA
Reese, L.C. (1997) Analysis of Laterally Loaded Piles in Weak Rock
(8)
GA, ID, MI, MT, NJ, NC, OR, TX
Carter and Kulhawy (1992) Analysis of Laterally Loaded Shafts in Rock
(1)
NJ
Other (please cite reference or provide a brief description)
ME: FB Pier evaluation
SC: Some lateral load resisting
WA: S-Shaft Program developed by M. Ashour and G. Norris of UNR along with J.P. Singh of
J.P. Singh & Associates. Model is based on strain wedge theory
23.

What, if any, computer programs are used by your agency for analysis of rock-socket response to lateral
loading?
LPILEPLUS
(23)
AL, AZ, CA, GA, HI, ID, IA, KY, MA, MI, MN, MT, NH, NJ, NM,
NC, OR, PR, SC, TX, UT, VT, WA
COM624P

(17) AR, CA, CT, GA, ID, IL, IA, KS, KY, ME, MA, NJ, NC, OR, PR, TX, VT

FBPIER

(8)

FL, MI, MN, NJ, NC, TN, TX, VT

Finite-Element Method (specify program)


NC: Flac

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

128

Other (provide name of program)


NH: Group 6.0
WA : S-Shaft Program developed by M. Ashour and G. Norris of UNR along with J.P. Singh of
J.P. Singh & Associates. Model is based on strain wedge theory
24.

If you use the p-y method of analysis, describe briefly how you determine the p-y relationships for rock
Published correlations between rock properties and p-y curve parameters
(4)
CA, KY, NJ, VT
Reference(s):

CA: LPILE Manual

Correlations built into computer code (specify program)


(22)
AZ, AR, CA, CT, FL, GA, HI, IL, IA, MA, MI, MN, NH, NM, NC, OR, PR, SC, TN,
TX, UT, VT
(all of the above states use LPILE)
In-house correlations based on agency experience
(2)
NC, OR
Educated guess
(2)
MA, OR
Other (describe)
CA: Pressuremeter Testing
MN: in situ test
NC: Research
OR: Limited pressuremeter data in soft rocks
WA: Reese p-y curves for vuggy limestone are derived using elastic theories. For basalt, using
engineering judgment, we typically define y as 0.01B, assuming 0.5% strain is the typical
range over which basalt behaves linearly and that is the rock within 2B that resists the load.
Therefore, y = 0.005(2B) or 0.01B. We then use correlations or published values to determine
Youngs modulus E. Typically, this is about 10,000 ksi for basalt. We then use the
unconfined compressive strength from point load tests along with E to define the curve. For
example, if qu is 22.8 ksi we would take the 10,000 ksi (E) value and divide by the qu to get
about 440. The p-y curve would then be defined by a straight line beginning at the origin with
a slope of 440 qu. In highly fractured rock, the engineer would use judgment to change strain
that defines y1, thus flattening the p-y curve.
25.

On projects completed by your agency, which of the following design considerations control rock-socket
length (approximately)?
Axial capacity
AL
AZ
CA
CT
HI
ID
IL
FL

Not really
99%
70%
50%
100%
100%
50%
70%

KS
KY
ME
MA
MI
MN
MT

50%
30%
100%
30%
50%
90%
50%

NJ
NM
NC
OR
SC
SD
TX
UT

65%
80%
30%
35%
40%
100%
95%
80%

KY:

60%

NC:

70%

Lateral load response


AL:

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

129

AZ:
AR:
CA:
CT:
GA:
IL:
KS:
FL:

1%
100%
20%
50%
10%
40%
50%
30%

MA:
MI:
MN:
MT:
NH:
NJ:
NM:

60%
50%
10%
50%
90%
35%
10%

OR:
SC:
TN:
TX:
UT:
VT:
WA :

60%
60%
5%
20%
100%
100%

IL:
KY:

10%
10%

MA:

10%

Construction-related
CA:
GA:

10%
90%

Sharing of load between drilled shaft and footing


Puerto Rico: 100%
Other (explain)
IA: No information available
KS: Minimum of 1.5 x shaft diameter
NM: 10% scour
OR: 5% scour
TN: 1.5 x socket diameter
26.

Please identify any other issues pertaining to rock-socket analysis/design that you feel should be addressed
in this synthesis.
FL: In karst, check for voids below tip.
IL: Bureau of Bridges and Structures, and consultants.
KS: Pertaining to Question 17, at the LFD load, the settlement should be less than some acceptable value.
We are using an arbitrary value of approximately 14 in. This will vary depending on the bridge type
and span length.
MA: The design/analysis of highway structures foundation (traffic signals, etc.). Construction practices
and QA/QC and their influences on design assumptions.
NH: Provide additional guidance for using side shear and end bearing in combination and provide
simplification of side resistance equations for cohesive soils contained in FHWA-IF-99-025.
OR: What agencies are using the AASHTO methods for drilled shaft design in rock?
UT: Concerns with appropriate lateral analysis methods; that is, is LPILE appropriate to be using with rock
sockets?

Part V: Structural Analysis


27.

What branch or group within your agency is responsible for structural design of rock-socketed drilled
shafts?
AL:
AZ:
AR:
CA:

Bridge Bureau
Bridge Group
Bridge Design
Division of Engineering Services/
Structure Design
CT: No drilled shaft design has been done
with in-house engineering staff
GA: Office of Bridge Design
HI: Bridge Design Section or Structural Consultants
ID: Bridge Design

ME: Bridge Program


MN: Bridge Office
MT: Bridge
NH: Bridge and Geotechnical Sections
NJ: Structural and Geotechnical Engineering Units
NM: Bridge Section
OR: Technical Services, Region Technical Centers
PR: None, done by consultants
SC: Bridge Design Section
SD: Bridge Design

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

130

IA: Office of Bridges and Structures


KS: Bridge Design Section
KY: Division of Bridge Design
MA: Bridge Designer (either in-house or consultant)
after consultation with Geotechnical Section

28.

TN: Division Structures


TX: Geotechnical Branch
UT: Structures Division
WA: Bridge and Structures Office
FL: Geotech for resistance; Structures for structure
design

Mark all of the applicable references/codes used by your agency in the structural design of rock-socketed
drilled shafts:
O'Neill and Reese (1999) Publication No. FHWA-IF-99-0 25, Drilled Shafts:
Construction Procedures and Design Methods
(15)
AR, CA, CT, FL, ID, IA, KS, KY, MA, NH, NJ, NC, OR, PR, VT
ACI 318, Building Code Requirements for Structural Concrete
(3)
IL, NJ, WA
AASHTO, Bridge Design Specifications
(26)
AZ, AR, CA, CT, FL, GA, HI, ID, IL, IA, KS, KY, MA, MN, NH, NJ, NM, NC, OR,
SC, SD, TN, TX, UT, VT, WA
ACI 336, Design and Construction of Drilled Piers
(2)
NJ, WA

29.

For structural design of drilled shafts, does your agency currently use Load Factor Design (LFD), Load and
Resistance Factor Design (LRFD), or Allowable Stress Design (ASD)?
SLD (allowable stress, or Service Load Design)
LRFD (Load and Resistance Factor Design)
LFD (Load Factor Design)
Mixed approach (SLD for foundation capacity and LFD or LRFD for load calculation)
SLD:

(7)

AZ, AR, GA, NM, NC, PR, TX

LRFD: (8)

CT, FL, HI, ID, ME, SC, UT, WA

LFD:

CA, KS, MA, KS, NC, WA

(6)

Mixed: (10)
(a) stated mixed only, no explanation: MN, NH, OR, SD, VT
(b) SLD for foundation capacity and LFD or LRFD for load calculation: KY, IL, IA, NJ, TN
30.

For structural design purposes, how would you best describe the analysis method used to obtain the
distribution of moment and shear with depth?
A point of fixity is assumed; shaft is then treated as a structural beam-column.
(11) CT, KS, KY, MA, NJ, NM, NC, SD, TN, TX, UT
Soil/Structure Interaction analysis is conducted using one of the following computer codes:
p-y method by COM624 or LPILEPlus
(21) AZ, AR, CA, GA, HI, ID, IL, IA, KS, ME, MA, MN, NJ, NM, OR, PR, SC, TX, UT, VT, WA
FBPIER

(5)

FL, ME, MN, NC, VT

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

131

Other
Elasticity solution

(2)

NH, WA (S-Shaft Program see above)

(1)

KS

Numerical methods such as finite element, boundary element, or finite difference specify
computer program:
none
Other method (explain briefly).
NM: Interaction with Geotechnical Section.
TN: Triangular stress distribution limited to side bearing capacity of rock and McCorkle side resistance
equations.
TX: Point of fixity is used for simple, typical structures. P-y method is used for more complex
structures.
31.

Please indicate whether you have encountered difficulties associated with the design or analysis issues
listed below and, if so, summarize the circumstances (shaft dimensions, depth of soil over rock, rock or
IGM type):
Unexpectedly high computed shear in the rock socket when using the p-y method of analysis.
CA: When the moments go from a maximum to zero over a relatively short length, then the
corresponding shear demands that are reported are large.
Difficulties or questions in applying p-y analysis to relatively short socket lengths
CA, IA, NM
NH: One question is whether the drilled shaft length can be terminated even though the p-y
analysis indicates some minor shear, moment, or deflection at the base of the shaft.
Questions regarding transfer of moment to the rock socket or development length
for reinforcing bars extending into the rock socket.
IA, MA

32.

Please identify any other issues pertaining to structural analysis/design that you feel should be addressed in
this synthesis.
MA: Should seismic design of rock-socket length be adequate to develop full plastic hinge moment
in reinforced concrete shaft?
OR: Not specifically related to rock sockets, but a design with about 6070 ft of overlying silt was
difficult to analyze. Resulting moments at superstructure were opposite direction of what would
be expected, did not tend to converge on a solution during seismic modeling runs. I chased it all
over the place (using LPILE, WinSTRUDL, ODOT BRIG2D software).

Part VI: Construction and Field Testing of Rock-Socketed Drilled Shafts


33.

Indicate whether the measures described below are included in construction specifications for rock or IGM
sockets designed by your agency:
Roughening of the sides of the socket by grooving or rifling
Restrictions on the use of slurry in sockets

(6)

AZ, IA, KS, ME, MA, MN

(14)

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

132

CA, FL (no polymer), GA, HI, KS, KY, ME, MA, MN, NM, NC, TX, UT, VT
Specifications for rock excavation by blasting (4)
34.

CA, ME, NC, OR

Does your agency specify requirements for cleanliness at the bottom of the excavation prior to
concrete placement?
Ye s

(28) AL, AZ, AR, CA, FL, GA, HI, ID, IL, IA, KS, KY, ME, MA, MN, MT, NH, NJ,
NM, NC, OR, SC, SD, TN, TX, UT, VT, WA

No

(1)

CT

If you answered Yes, please provide a brief summary of the following:


Requirements for cleanliness:
Six states (FL, HI, IL, NH, NC, and SC) gave the following: minimum 50% of base area to have
less than 0.5 in. and maximum depth not to exceed 1.5 in.
AR: No more than 1 in. of loose material.
CA: Specification simply states that the contractor verifies that the bottom is clean.
CT: Not written into specification, but generally following recommendations in FHWA Drilled Shaft
Manual.
GA: No loose sediment or debris.
ID: Less than 2 in. thick for end bearing shafts; less than 6 in. for side friction shafts.
IA: Minimum 50% of base to have less than 0.5 in. and maximum depth not to exceed 1 in.
KS: Just prior to placing concrete, a minimum of 75% of the base must have less than 0.5 in. of
sediment; CSL also used for wet pours.
KY: Maximum 0.5 in. of sediment.
MA: End bearing <1 in., skin friction <3 in.
ME: Minimum of 50% of the base of shaft should have <0.5 in. of sediment at time of concrete
placement.
MN: From our drilled shaft special provision: loose material shall be removed from drilled shafts
prior to placement of reinforcement. After the shafts have been cleaned, the engineer will inspect the
shafts for conformance to plan dimensions and construction tolerances. If permanent casing is
damaged and unacceptable for inclusion in the finished shaft, the casing shall be replaced at the
contractors expense. If a portion of a shaft is underwater, the contractor shall demonstrate that the
shaft is clean to the satisfaction of the Engineer. This shall include inspection by a diver, at no cost to
the department, if considered necessary by the Engineer. Dewatering of the drilled shafts for
cleaning, inspection, and placement of reinforcement and concrete will not be required. If the drilled
shaft contractor chooses to dewater the shafts for convenience of construction, this work shall be done
at the contractors expense.
NJ: Less than 0.5 in. of sediment.
NM: <1 in. of loose material.
OR: No more than 2 in. of loose material for end-bearing; no more than 6 in. of loose material for
friction shafts. Assume end-bearing if not specified.
SD: Make sure the bottom of the shaft is free of loose material.
TN: No loose soil or rock cuttings allowed.
TX: From Specification 416 Drilled Shaftsremove loose material and accumulated seep water
from the bottom of the excavation prior to placing the concrete.
UT: Remove all lose material from the bottom of drilled holes before placing concrete.
WA: The contractor shall use appropriate means such as a cleanout bucket or air lift to clean the
bottom of the excavation of all shafts. No more than 2 in. of loose or disturbed material shall be
present at the bottom of the shaft just prior to placing concrete for end bearing shafts. No more than 6
in. of loose or disturbed material shall be present for side friction shafts. End bearing shafts shall be
assumed unless otherwise noted in the contract. Shafts specified as both side friction and end bearing
shall conform to the sloughing criteria specified for end bearing shafts.

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

133

Method(s) and tools used to verify cleanliness requirement:


AZ: Visual and hand probe.
AR: Video equipment or person in hole with suitable lighting and ventilation.
CA: Contractor submits a concrete placement plan for approval. Usually, they will use a clean out
bucket to clean the bottom of hole. There is inspection of the drilling slurry at bottom of the shaft
prior to placing concrete. Caltrans occasionally will verify the bottom of end bearing shafts with a
camera.
CT: Probing of the rock-socket bottom.
FL: Probing, sometimes use SID.
GA: Hand cleaned and inspected.
HI: Weighted tape.
ID: Cleaning buckets or air lift.
IA: Weighted tape, camera inspection (rare), sediment deposition trial run in open-top clean-out
bucket.
KS: Visual on dry pours, sounding (using probes) underwater.
KY: Judgment of inspector.
MA: Weighted rods, visual check by use of cleaning equipment spoils.
ME: Weighted tape and remotely operated cameras.
MN: See above.
NH: Weighted tape or solid rod.
NJ: Sounding by weighted tape.
NM: Weighted tape/sounding.
NC: 1. Visual Inspection. 2. Steel Probe (10 lb weight). 3. SID shaft inspection device.
OR: Weighted tape, rod probe or visual.
SD: Visual inspection if the shaft is dry, otherwise use a clean-out bucket.
TN: Visual inspection.
TX: Visual, clean-out bucket.
UT: Visual.
WA: The excavated shaft shall be inspected and approved by the Engineer prior to proceeding with
construction. The bottom of the excavated shaft shall be sounded with an airlift pipe, a tape with a
heavy weight attached to the end of the tape, or other means acceptable to the engineer to determine
that the shaft bottom meets the requirements in the contract.
35.

Does your agency use construction specifications or special provisions that account for construction of
Ye s
No
sockets in a particular rock type?
Yes:

(3) AZ, ID, KY

No:

(25)

If Yes, please provide a brief description.


Rock type and special provision:
AZ: Limestone; drill below tip elevation to check for karst conditions.
ID: Special provisions for IGMs and hard rock.
KY: Soft shale; sometimes require contractors to use polymer slurry in wet holes with soft shales
subject to slaking in the presence of water.
CA: Different pay items for Cast-in-Drilled-Hole concrete piling and Cast-in-Drilled-Hole (Rock
Socket) concrete piling.
36.

Have you observed any methods, equipment, or materials used for socket construction that you believe are
a source of construction problems?
Ye s

No

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

134

Yes :

(10)

CA, FL, KS, KY, NH, NC, SD, TN, UT, VT

No:

(16)

AL, AZ, AR, CT, GA, HI, ID, IL, IA, ME, MA, MN, OR, SC, TX, WA

If yes, please explain:


CA: We design for low bidding contractors to get the contract and the construction problems that will
result. Rock may be harder than the contractor thought when bidding and planning the job. Thus, the
drilling equipment brought out is often unable to drill or very slow to drill the rock. This results in
costly contractor claims.
FL: Improper equipment or size; full-length casing reduces skin friction.
KS: In wet pours, inadequate sealed tremie or no pig in the concrete pump supply line. Loss of
slump in the concrete during placement. Dirty bottoms were observed with Sonic testing.
KY: Reverse circulation drilling methods used in conjunction with polymer slurry when used as
described in Question 35.
NH: Certain clean-out buckets cannot always meet the cleanliness requirements.
NC: Various methods used to force a dry pour.
SD: If not done properly using pump trucks to place concrete can cause soft bottoms in the shafts.
TN: We have significant depths of interlayered rock layers and soil that makes it difficult to use
either an auger or a core barrel.
UT: Concerns with use of drilling fluids instead of casing.
VT: In holes cased through the overburden soils into bedrock we have had problems seating the
casing into rock. This was true when the contractor used a casing diameter that was essentially the
same as the rock-socket diameter. The casing was vibrated into the rock-socket hole, which resulted
in more rock drilling than expected, because the casing followed the socket. This resulted in longer
overall shaft lengths than planned, particularly when the upper portion of the rock was fractured or
weathered.
37.

Please identify any other construction-related issues for rock or IGM sockets that you believe should be
addressed in this synthesis.
FL: What is rock, where does it start, quality.
KS: Pertaining to Question 35Our special provision accounts for a wet or dry pour (cased or
uncased) rather than rock type.
MA: Define Top of Rock, which generally can be a discrepancy between borings and
construction drilling.
NH: Effect of slurry on side friction.
OR: If during construction the top of rock elevation is found to be different than what was
assumed in design, what is the effect? How different does it have to be to have a significant effect
on design?

38.

Indicate whether your agency has used any of the following field load testing methods on rock-socketed
drilled shafts.
Conventional static axial load test

(7)

CA, FL, GA, IA, NC, TX, UT

Conventional lateral load test

(5)

CA, FL, MA, NJ, NC

Osterberg Cell for axial load test

(18)

CA, CT, FL, GA, HI, IL, IA, KS, KY, ME,
MA, MN, NJ, NM, NC, PR, SC, TX

Osterberg Cell for lateral load test

(1)

SC

Statnamic test for axial load

(6)

CT, FL, IA, NC, PR, SC

Statnamic test for lateral load

(3)

AL, FL, SC

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

135

High strain impact

(3)

FL, KS, MA

If your agency has used the Osterberg Cell (O-cell) for axial load tests on rock-socketed shafts, please answer the
following:
39.

Were you able to measure both side and tip resistances of the socket independently?
Ye s
Yes:
No:

40.

No
(17) CA, CT, FL, GA, HI, IL, IA, KS, KY, ME, MA, NJ, NM, NC, PR, SC, TX
(1) MN

Was the test used to determine


Ultimate side resistance (of socket)
(14) CA, CT, FL, GA, IA, KS, KY, ME, MA, MN, NM, NC, SC, TX
Ultimate base resistance
(8) IL, FL, IA, KS, KY, NC, SC, TX
Proof load only
(4) HI, NJ, PR, SC

41.

Additional comments regarding use of O-cell for load testing of rock sockets.
IL: Too expensive.
IA: None.
KY: We have typically failed shafts in side resistance and mobilized enough base movement to extrapolate
the ultimate base resistance.
ME: Did not mobilize ultimate base resistance on either project.
MN: Was also used to develop p-y curve.
NM: Use was for IGM Santa Fe Formation.
TX: For information on recent O-cell testing in rock contact Dr. Vipu and University of Houston.

If your agency has experience with Statnamic testing of rock-socketed drilled shafts please answer the following:
42.

Which of the following performance parameters were determined by the test? (Check all that apply.)
Socket side resistance
(5) FL, AL, IA, NC, SC
Total socket resistance (side and base)
(4) FL, NC, PR, SC
Lateral load-displacement response
(3) FL, NC, SC

43.

Socket base resistance


(5) CT, FL, IA, NC, SC
Axial load displacement response
(5) AL, FL, NC, PR, SC

Additional comments regarding Statnamic testing of rock-socketed drilled shafts.


FL: Limit on size that can be tested.
NC: Test is very expensive; we need to find another method with less cost.

44.

Do you have results of load tests on rock-socketed drilled shafts and, if so, are you willing to receive
follow-up contact regarding the possibility of using your results for the synthesis?
Yes, I have previously unpublished load test results
(9) CA, GA, IA, KS, KY, ME, MA, NC, SC

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

136

Yes, I am willing to receive follow-up contact


(11) CA, CT, GA, IA, KS, KY, ME, MA, NM, NC, PR
If previously published, please give a reference:
ME: Loadtest, Inc., performed all O-cell tests on the BathWoolwich and HancockSullivan
bridges in Maine.
MN: Transportation Research Record 1633.
45.

Indicate which of the following nondestructive testing methods are used on a regular basis by your agency
for rock-socketed shafts.
gamma-gamma
crosshole sonic logging
sonic echo
impulse response
parallel seismic
other

46.

(3) AZ, AR, CA


(20) AZ, AR, CA, CT, HI, ID, IA, KS, KY, ME, MA, MN, NJ, NM,
NC, OR, PR, SC, SD, VT
(1) AR
(1) AR
(3) CA (downhole camera), NC (Osterberg load cell), VT (CSLT
one project)

Based on your experience, are there any special considerations or issues related to the use of NDT-NDE,
specifically for rock-socketed shafts? If so, explain.
FL: Results are iffy.
IA: None.
KY: No.
MA: The configuration of the test pipes within the socket (if diameter is smaller than shaft) and
the possible influence of rock material properties on the data results.
NM: Sonic echo not utilized.
NC: Technology is not 100% accurate.
PR: We bought the equipment (CSL) last month.

47.

Do you have case histories of design, construction, or testing of rock-socketed drilled shafts that, in your
opinion, could provide useful information to your colleagues and, if so, are you willing to be contacted by
the author of the synthesis to discuss your case histories further?
Yes, I have useful case histories

(9) CA, CT, IA, KS, KY, ME, NM, NC, WA

Yes, I am willing to receive follow-up contact

(8) CA, CT, GA, IA, KS, KY, ME, NM

Copyright National Academy of Sciences. All rights reserved.

Rock-Socketed Shafts for Highway Structure Foundations

Abbreviations used without definitions in TRB publications:


AASHO
AASHTO
ACRP
ADA
APTA
ASCE
ASME
ASTM
ATA
CTAA
CTBSSP
DHS
DOE
EPA
FAA
FHWA
FMCSA
FRA
FTA
IEEE
ISTEA
ITE
NASA
NCFRP
NCHRP
NHTSA
NTSB
SAE
SAFETEA-LU
TCRP
TEA-21
TRB
TSA
U.S.DOT

American Association of State Highway Officials


American Association of State Highway and Transportation Officials
Airport Cooperative Research Program
Americans with Disabilities Act
American Public Transportation Association
American Society of Civil Engineers
American Society of Mechanical Engineers
American Society for Testing and Materials
American Trucking Associations
Community Transportation Association of America
Commercial Truck and Bus Safety Synthesis Program
Department of Homeland Security
Department of Energy
Environmental Protection Agency
Federal Aviation Administration
Federal Highway Administration
Federal Motor Carrier Safety Administration
Federal Railroad Administration
Federal Transit Administration
Institute of Electrical and Electronics Engineers
Intermodal Surface Transportation Efficiency Act of 1991
Institute of Transportation Engineers
National Aeronautics and Space Administration
National Cooperative Freight Research Program
National Cooperative Highway Research Program
National Highway Traffic Safety Administration
National Transportation Safety Board
Society of Automotive Engineers
Safe, Accountable, Flexible, Efficient Transportation Equity Act:
A Legacy for Users (2005)
Transit Cooperative Research Program
Transportation Equity Act for the 21st Century (1998)
Transportation Research Board
Transportation Security Administration
United States Department of Transportation

Copyright National Academy of Sciences. All rights reserved.

Anda mungkin juga menyukai